United States
Environmental Protection
M lkAgency
EPA/600/R-18/097
June 2018
www.epa.gov/isa
Integrated Science Assessment
for Oxides of Nitrogen, Oxides of
Sulfur, and Particulate Matter—
Ecological Criteria
(Second External Review Draft)
National Center for Environmental Assessment-RTP Division
Office of Research and Development
U.S. Environmental Protection Agency
Research Triangle Park, NC
June 2018
i
DRAFT: Do Not Cite or Quote

-------
Disclaimer
This document is an external review draft, for review purposes only. This information is
distributed solely for predissemination peer review under applicable information quality
guidelines. It has not been formally disseminated by EPA. It does not represent and
should not be construed to represent any Agency determination or policy. Mention of
trade names or commercial products does not constitute endorsement or recommendation
for use.
June 2018
ii
DRAFT: Do Not Cite or Quote

-------
CONTENTS
INTEGRATED SCIENCE ASSESSMENT TEAM FOR OXIDES OF NITROGEN, OXIDES
OF SULFUR, AND PARTICULATE MATTER—ECOLOGICAL CRITERIA _XXXII
AUTHORS, CONTRIBUTORS, AND REVIEWERS	XXXIV
CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE OXIDES OF NITROGEN, OXIDES OF
SULFUR, AND PARTICULATE MATTER—ECOLOGICAL CRITERIA
NAAQS REVIEW PANEL	XXXIX
ACRONYMS AND ABBREVIATIONS	XLI
PREFACE		LI
Legislative Requirements for the Review of the National Ambient Air Quality Standards	li
Overview and History of the Reviews of the Secondary National Ambient Air Quality Standards
for Nitrogen Dioxide, Sulfur Dioxide, and Particulate Matter	 lii
Nitrogen Dioxide Secondary National Ambient Air Quality Standards	 lii
Sulfur Dioxide Secondary National Ambient Air Quality Standards	liii
Particulate Matter Secondary National Ambient Air Quality Standards	Ivii
Most Recent Combined Review of the Oxides of Nitrogen and Oxides of Sulfur
National Ambient Air Quality Standards	 Ix
EXECUTIVE SUMMARY	LXIII
Purpose and Scope of the Integrated Science Assessment	Ixiii
Figure ES-1 Roadmap of the Integrated Science Assessment linking
atmospheric concentrations and deposition, soil and aquatic
biogeochemistry, and biological effects. 	Ixvi
Emissions, Ambient Air Concentrations, and Deposition	Ixvi
Ecological Effects	Ixviii
Table ES-1 Causal determinations for relationships between criteria
pollutants and ecological effects from the 2008 NOx/SOx ISA
or the 2009 ISA for particulate matter (PM), for other effects of
PM, and the current draft Integrated Science Assessment.	Ixix
Figure ES-2 Causal relationships between the criteria pollutants and
ecological effects. 	Ixxiii
Direct Phytotoxic Effects of Gas-Phase Oxides of Nitrogen (NOy) and Oxides of
Sulfur (SOx)	Ixxiv
Ecological Effects of Nitrogen and Sulfur Deposition 	Ixxiv
Ecosystem Services 	Ixxix
Integrating across Ecosystems	Ixxx
Figure ES-3 Causal relationships between the criteria pollutants and
ecological effects organized under ecosystem type.	Ixxxi
CHAPTER 1 INTEGRATED SYNTHESIS	1-1
1.1 Introduction to This Integrated Science Assessment (ISA)	1-1
1.1.1	Purpose	1-1
1.1.2	Process and Development	1-4
Figure 1-1 Workflow for collecting relevant literature for the 2017
Integrated Science Assessment for Oxides of Nitrogen, Oxides
of Sulfur, and Particulate Matter—Ecological Criteria.	1-5
1.1.3	Organization	1-6
June 2018
iii
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
1.2	Connections, Concepts, and Changes	1-7
1.2.1	Connections	1-7
Figure 1-2 Overview of atmospheric chemistry, deposition, and ecological
effects of emissions of oxides of nitrogen, oxides of sulfur, and
reduced nitrogen.	1-8
1.2.2	Concepts	1-8
Figure 1-3 An example of the matrix of information considered in defining
and calculating critical loads (see discussion in text). Note that
multiple alternative biological indicators, critical biological
responses, chemical indicators, and critical chemical limits
could be used.	1-13
1.2.3	Changes: New Evidence and Causal Determinations	1-20
Table 1-1 Causal determinations for relationships between criteria
pollutants and ecological effects from the 2008 NOx/SOx ISA
or the 2009 PM ISA, for other effects of particulate matter
(PM), and the current draft Integrated Science Assessment. 	1-22
Figure 1-4 Causal relationships between the criteria pollutants and
ecological effects. 	1-25
1.3	Emissions and Atmospheric Chemistry	1-27
1.3.1	Sources and Atmospheric Transformations	1-28
1.3.2	Measurement and Modeling Techniques	1-28
1.3.3	Spatial and Temporal Variability in Deposition	1-29
Figure 1 5 Three-year (2011 to 2013) average annual dry + wet
deposition of total oxidized nitrogen and reduced nitrogen
species in kilograms of nitrogen per hectare per year. 	1-31
Figure 1-6 Total deposition of total oxidized nitrogen, reduced nitrogen,
and oxidized sulfur expressed as H+ equivalents per hectare
per year over the contiguous U.S. 2011-2013.	1-32
1.4	Gas-Phase Direct Phytotoxic Effects	1-33
1.4.1	Sulfur Dioxide	1-33
1.4.2	Nitrogen Oxide, Nitrogen Dioxide, and Peroxyacetyl Nitrate	1-34
1.4.3	Nitric Acid	1-34
1.5	Terrestrial Ecosystem Nitrogen Enrichment and Acidification	1-34
1.5.1	Soil Biogeochemistry	1-36
Table 1-2 Summary of key soil geochemical processes and indicators
associated with eutrophication and acidification.	1-37
1.5.2	Biological Effects of Terrestrial Nitrogen Enrichment	1-42
Figure 1-7 Summary of critical loads in the U.S. for shrubs and
herbaceous plants (yellow), trees (blue), lichens (green), and
mycorrhizae (grey).	1-52
1.5.3	Biological Effects of Acidification	1-53
Figure 1-8 Forest ecosystem critical loads for soil acidity related to base
cation soil indicators.	1-56
1.6	Freshwater Ecosystem Nitrogen Enrichment and Acidification	1-57
1.6.1	Freshwater Biogeochemistry	1-58
Table 1-3 Summary of key aquatic geochemical processes and indicators
associated with eutrophication and acidification.	1-60
1.6.2	Biological Effects of Freshwater Nitrogen Enrichment	1-65
1.6.3	Biological Effects of Freshwater Acidification	1-69
Figure 1-9 Surface water critical loads for acidity in the U.S. 10th
percentile aggregation for36-km2 grids with S and N.	1-74
1.7	Estuarine and Near-Coastal Ecosystem Nitrogen Enrichment and Acidification	1-76
1.7.1	Estuary and Near-Coastal Biogeochemistry	1-76
1.7.2	Biological Effects of Nitrogen Enrichment	1-80
1.7.3	Biological Effects of Nutrient-Enhanced Coastal Acidification	1-83
June 2018
iv
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
1.7.4 National-Scale Sensitivity and Critical Loads	1-83
1.8	Wetland Ecosystem Nitrogen Enrichment and Acidification	1-85
1.8.1	Wetland Biogeochemistry	1-86
1.8.2	Biological Effects of Wetland Nitrogen Enrichment/Eutrophication	1-87
1.9	Freshwater and Wetland Ecosystem Sulfur Enrichment	1-90
1.9.1	Biogeochemistry	1-91
1.9.2	Biological Effects of Sulfur Enrichment	1-92
1.9.3	National-Scale Sensitivity and Critical Loads	1-95
1.10	Ecological Effects of Particulate Matter other than N and S Deposition	1-95
1.11	Recovery of Ecosystems from Nitrogen (N) and Sulfur (S) Deposition in the U.S.	1-97
1.11.1	Overarching Concepts of Ecological Recovery from Acidification	1-97
1.11.2	Acidification Recovery in the U.S.	1-99
1.11.3	Nitrogen (N)-Driven Nutrient Enrichment Recovery in the U.S.	1-100
1.12	Climate Modification of Ecosystem Response to Nitrogen (N) and Sulfur (S) Addition	1-100
1.13	Key Scientific Uncertainties 	1-101
1.13.1	Atmospheric Science 	1-102
1.13.2	Ecological Effects	1-105
1.14	Ecosystem Services	1-110
APPENDIX 1. INTRODUCTION TO APPENDICES	1-1
APPENDIX 2. SOURCE TO DEPOSITION	2-1
2.1.	Introduction	2-1
2.2.	Sources of Nitrogen and Sulfur Compounds and Particulate Matter to the Atmosphere 	2-5
2.2.1.	National Emissions by Source	2-6
Table 2-1 Emissions of NOx (nitric oxide + nitrogen dioxide), sulfur
dioxide, and ammonia by source category for 2014
(Teragrams3 N, S/yr).	2-7
2.2.2.	Methods of Estimating Emissions	2-8
Figure 2-1 Modeling system used to compute 2014 Fertilizer Application
Emissions.	2-11
Figure 2-2 Process to produce specific location and practice specific daily
emission factors for livestock waste.	2-12
2.2.3.	Evaluation and Uncertainty	2-13
2.3.	Atmospheric Chemistry of Nitrogen and Sulfur Species and Particulate Matter (PM)	2-16
2.3.1.	Nitrogen Oxides 	2-16
Figure 2-3 Schematic diagram showing pathways for reactive nitrogen
species in ambient air.	2-17
Table 2-2 Henry's law coefficients for selected reactive nitrogen species
at 25°C in water.	2-18
2.3.2.	Sulfur Oxides	2-20
Figure 2-4 Rate of conversion of sulfur (IV) to sulfur (VI) by different
oxidation paths as a function of pH.	2-21
2.3.3.	Acid Neutralization by Ammonia	2-22
2.3.4.	Organic Nitrogen and Sulfur	2-24
2.3.5.	Organic Acids	2-25
2.3.6.	Particulate Matter (PM)	2-26
Figure 2-5 Contributions of organic carbon (OC), elemental carbon (EC),
sulfate, nitrate, sea salt, and crustal components to PM2.5 at
selected sites (A) 2003-2005 (B) 2013-2015. 	2-27
2.4.	Concentration and Deposition Measurements	2-28
2.4.1. Monitoring Networks	2-28
Table 2-3 Summary of monitoring networks used by Schwede and Lear
(2014a).	2-30
June 2018
v
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
2.4.2.	NO2, NOx, and NOy 	2-32
2.4.3.	Ammonia	2-34
2.4.4.	Sulfur Dioxide	2-37
Table 2-4 Sources of uncertainty for individual Ozone Monitoring
Instrument measurements in the study of Nowlan et al. (2014).	2-38
2.4.5.	Filter-Based Concentration Measurements	2-38
Figure 2-6 Clean Air Status and Trends Network filter pack. 	2-39
Figure 2-7 Comparison between weekly average measurements of sulfur
dioxide using the Clean Air Status and Trends Network filter
pack and the trace ultraviolet pulsed fluorescence monitor in
2014.	2-41
2.4.6.	Deposition Measurements 	2-42
2.5.	Modeling Chemistry, Transport, and Deposition	2-45
2.5.1.	Advances in Chemistry-Transport Model (CTM) Modeling	2-45
2.5.2.	Modeling Deposition	2-47
Figure 2-8 Schematic diagram showing mechanisms for transferring
pollutants from the atmosphere to the surface.	2-48
Table 2-5 Average dry deposition velocities (cm/s) for several gases over
land surfaces. 	2-49
Table 2-6 Deposition velocity (cm/s) for sulfur dioxide averaged over
different land use types for summer and winter.	2-49
Figure 2-9 Modeled and measured deposition velocities over grass (left
figure) and coniferous forest canopies (right figure) for particles
of density 1 gm/cm3 depositing under similar friction velocity
(u*) (35 
-------
CONTENTS (Continued)
Figure 2-17 Wet deposition of ammonium + nitrate (kg N/ha/yr) over the
contiguous U.S. in two, 3-year periods, 2012 to 2014 and 1989
to 1991. Also shown are active National Trends Network sites
in either period.	2-71
2.6.3.	Oxidized Nitrogen	2-72
Figure 2-18 Distribution of annual NOx emissions in 2001. 	2-73
Figure 2-19 Distribution of annual average total oxidized nitrogen species
concentrations for 2011 simulated by Community Multiscale Air
Quality modeling system.	2-74
Figure 2-20 Seasonal average surface nitrogen dioxide mixing ratios in
parts per billion for winter (upper panel) and summer (lower
panel) derived by the Ozone Monitoring
Instrument/GEOS-Chem model for 2009-2011. The Ozone
Monitoring Instrument has an overpass at approximately
1:30 p.m. local standard time.	2-75
Figure 2-21 Three-year average (2011 -2013) surface concentrations of
nitric acid based on monitoring data obtained at Clear Air
Status and Trends Network sites (black dots). 	2-77
Figure 2-22 Three-year average (2011 -2013) surface concentrations of
particulate nitrate based on monitoring data obtained at Clear
Air Status and Trends Network sites (black dots).	2-78
Figure 2-23 Total oxidized nitrogen deposition over the contiguous U.S.
2011 -2013.	2-79
Figure 2-24 (Left) nitrate wet deposition, 1989-1991; (Right) nitrate wet
deposition, 2012-2014.	2-80
Figure 2-25 Difference in wet deposition of nitrate (kg N/ha/yr) over the
contiguous U.S. between 1989 to 1991 and 2012 to 2014. The
range of positive values is smaller than that for negative
values.	2-81
2.6.4.	Reduced Nitrogen	2-81
Figure 2-26 Distribution of annual ammonia (NH3) emissions by county in
2001.	2-83
Figure 2-27 Average (2012) surface concentration of ammonia obtained by
the Ambient Ammonia Monitoring Network at select Clear Air
Status and Trends Network sites. Concentrations of ammonia
(|jg/m3) can be converted to mixing ratios (parts per billion) to
rough approximation at normal temperature and pressure by
multiplying by 1.4.	2-84
Figure 2-28 Three-year average (2011 -2013) surface concentrations of
particulate ammonium (|jg/m3) based on monitoring data
obtained at Clear Air Status and Trends Network sites (black
dots).	2-85
Figure 2-29 Total reduced inorganic nitrogen deposition over the
contiguous U.S. 2011 -2013.	2-86
Figure 2-30 (Left) ammonium wet deposition, 1989-1991; (Right)
ammonium wet deposition, 2012-2014.	2-87
Figure 2-31 Difference in wet deposition of ammonium (kg N/ha/yr) over
the contiguous U.S. between 1989 to 1991 and 2012 to 2014.	2-88
2.6.5.	Sulfur Oxides	2-88
Figure 2-32 Distribution of annual sulfur dioxide (SO2) emissions by county
from the 2014 National Emissions Inventory. 	2-89
Figure 2-33 Three-year average (2011 -2013) surface concentrations of
sulfur dioxide obtained by fusion of monitoring data obtained at
Clear Air Status and Trends Network sites (black dots) and
Community Multiscale Air Quality model system results.
Concentrations (|jg/m3) can be converted to mixing ratios
(parts per billion) at normal temperature and pressure) to rough
approximation by multiplying by 0.37.	2-90
June 2018	vii	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 2-34 Three-year average (2011 -2013) surface concentrations of
particulate sulfate based on monitoring data obtained at Clean
Air Status and Trends Network sites (black dots).	2-91
Figure 2-35 Total deposition of sulfur (kg S/ha/yr) over the contiguous U.S.
2011 -2013.	2-92
Figure 2-36 Percentage of deposition of total sulfur as dry deposition over
the contiguous U.S. 2011-2013.	2-93
Figure 2-37 (Left) sulfate wet deposition, 1989-1991; (Right) sulfate wet
deposition, 2012-2014.	2-94
Figure 2-38 Difference in wet deposition of sulfate (kg S/ha/yr) over the
contiguous U.S. between 1989 to 1991 and 2012 to 2014. The
range of positive values is much smaller than for negative
values.	2-94
2.6.6.	Particulate Matter (PM)	2-95
Figure 2-39 3-year average concentrations of particulate matter smaller
than 2.5 |jm diameter (PM2.5) 2013-2015.	2-96
Figure 2-40 National monthly average concentrations of particulate matter
smaller than 2.5 |jm diameter (PM2.5; top) and sulfate in PM2.5
(bottom) from 2000-2015 (concentrations in |jg/m3). 	2-97
Figure 2-41 98th percentile concentrations for PM10-2.5 between
2013-2015.	2-98
2.6.7.	Distributions of Dry Deposition of Nitrogen Dioxide and Sulfur Dioxide
Derived Using Satellite-Based Measurements and Chemistry Transport
Models 	2-98
Figure 2-42 Top panel: modeled deposition velocities for nitrogen dioxide
and sulfur dioxide for 2005 to 2007; middle panel
Satellite-model estimates of annual mean dry deposition fluxes
of nitrogen dioxide and sulfur dioxide; bottom panel:
uncertainties in estimates.	2-100
2.6.8.	Background Concentrations and Deposition	2-101
Figure 2-43 Contributions to oxidized and reduced nitrogen deposition from
U.S.: anthropogenic (top), foreign anthropogenic (middle), and
natural sources (bottom).	2-102
2.7. Supplemental Material on Changes in Deposition since 2000 	2-105
Figure 2-44 Wet plus dry deposition of total nitrogen over 3-year periods.
Top: 2000-2002; Bottom: 2011-2013. 	2-107
Figure 2-45 Wet deposition of total nitrogen over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-108
Figure 2-46 Dry deposition of total nitrogen over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-109
Figure 2-47 Percent of total nitrogen as dry deposition over 3-year periods.
Top: 2000-2002; Bottom: 2011-2013. 	2-110
Figure 2-48 Wet plus dry deposition of oxidized nitrogen over 3-year
periods. Top: 2000-2002; Bottom: 2011-2013. 	2-111
Figure 2-49 Percent of total nitrogen as oxidized nitrogen over 3-year
periods. Top: 2000-2002; Bottom: 2011-2013. 	2-112
Figure 2-50 Dry deposition of oxidized nitrogen over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-113
Figure 2-51 Percent of total nitrogen dry deposited as oxidized nitrogen
over 3-year periods. Top: 2000-2002; Bottom: 2011-2013.	2-114
Figure 2-52 Combined dry deposition of nitric acid and particulate nitrate
over 3-year periods. Top: 2000-2002; Bottom: 2011-2013.	2-115
Figure 2-53 Dry deposition of nitric acid over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-116
Figure 2-54 Dry deposition of particulate nitrate over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-117
June 2018
viii
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 2-55 Dry deposition of modeled (unmeasured) nitrogen species over
3-year periods. Top: 2000-2002; Bottom: 2011-2013.	2-118
Figure 2-56 Percent of total nitrogen as modeled (unmeasured) species
over 3-year periods. Top: 2000-2002; Bottom: 2011-2013.	2-119
Figure 2-57 Wet plus dry deposition of reduced (inorganic) nitrogen over
3-year periods. Top: 2000-2002; Bottom: 2011-2013.	2-120
Figure 2-58 Percent of total nitrogen deposition by reduced (inorganic)
nitrogen over 3-year periods. Top: 2000-2002; Bottom:
2011-2013.	2-121
Figure 2-59 Dry deposition of ammonia over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-122
Figure 2-60 Dry deposition of particulate ammonium over 3-year periods.
Top: 2000-2002; Bottom: 2011-2013. 	2-123
Figure 2-61 Dry deposition of reduced (inorganic) nitrogen over 3-year
periods. Top: 2000-2002; Bottom: 2011-2013. 	2-124
Figure 2-62 Percent of total nitrogen deposition by dry reduced (inorganic)
nitrogen over 3-year periods. Top: 2000-2002; Bottom:
2011 -2013.	2-125
Figure 2-63 Wet plus dry deposition of total sulfur over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-126
Figure 2-64 Wet deposition of total sulfur over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-127
Figure 2-65 Dry deposition of total sulfur over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-128
Figure 2-66 Percent of total sulfur deposition by dry deposition over 3-year
periods. Top: 2000-2002; Bottom: 2011-2013. 	2-129
Figure 2-67 Dry deposition of sulfur dioxide over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-130
Figure 2-68 Dry deposition of particulate sulfate over 3-year periods. Top:
2000-2002; Bottom: 2011-2013. 	2-131
APPENDIX 3. DIRECT PHYTOTOXIC EFFECTS OF GASEOUS OXIDIZED NITROGEN
AND SULFUR ON VEGETATION 	3-1
3.1.	Introduction	3-1
Figure 3-1 The microarchitecture of a dicot leaf. While details among
species vary, the general overview remains the same. Light
that drives photosynthesis generally falls upon the upper
(adaxial) leaf surface. Carbon dioxide, oxides of sulfur, oxides
of nitrogen, and ozone gases generally enter by diffusion
through the guard cells (or stomata) on the lower (abaxial) leaf
surface, while water vapor exits through the stomata
(transpiration).	3-3
3.2.	Direct Phytotoxic Effects of Sulfur Dioxide on Vegetation	3-3
Figure 3-2 Map of maximum 3-hour daily max average sulfur dioxide
concentration reported at Air Quality System monitoring sites
for 2016. 	3-5
3.3.	Direct Phytotoxic Effects of Nitric Oxide, Nitrogen Dioxide, and Peroxyacetyl Nitrate	3-7
Figure 3-3 Map of U.S. annual average nitrogen dioxide concentrations
for 2013. 	3-8
3.4.	Direct Phytotoxic Effects of Nitric Acid	3-13
3.5.	Summary	3-15
3.5.1.	Sulfur Dioxide	3-15
3.5.2.	Nitrogen Oxide, Nitrogen Dioxide, and Peroxyacetyl Nitrate	3-16
3.5.3.	Nitric Acid	3-16
APPENDIX 4. SOIL BIOGEOCHEMISTRY	4-1
June 2018	ix	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
4.1.	Introduction	4-1
4.2.	Nitrogen and Sulfur Sources to Soil	4-2
Figure 4-1 Dominant sources of nitrogen across the U.S. at 8-digit
hydrologic unit codes.	4-3
Figure 4-2 Percentage of nitrogen input from nitrogen deposition at 8-digit
hydrologic unit codes.	4-3
4.3.	Soil Pools and Processes	4-4
Table 4-1 Summary of key soil geochemical processes and indicators
associated with eutrophication and acidification.	4-6
4.3.1.	Nitrogen Pathways and Pools	4-6
Table 4-2 Pathways and pools.	4-8
4.3.2.	Nitrogen Accumulation, Saturation, and Leaching 	4-11
Figure 4-3 A conceptual framework for the responses of the ecosystem
nitrogen (N) cycle to nitrogen (N) addition. 	4-12
Figure 4-4 A hypothetical model to account for the effects of nitrogen
supply on plant nitrogen uptake and belowground carbon
allocation, microbial production of inorganic nitrogen, and
nitrogen leaching. 	4-15
Table 4-3 Nitrogen accumulation, saturation, and leaching.	4-16
4.3.3.	Sulfate Accumulation, Adsorption, and Leaching	4-22
Table 4-4 Sulfate adsorption, accumulation and leaching. 	4-24
4.3.4.	Base Cation Leaching and Exchange	4-27
Table 4-5 Base cation leaching and exchange.	4-30
4.3.5.	Aluminum Mobilization	4-35
Table 4-6 Aluminum mobilization.	4-37
4.3.6.	Nitrification and Denitrification	4-39
Table 4-7 Nitrification and denitrification.	4-41
Figure 4-5 Effects of nitrogen addition on biogenic nitrous oxide emission.	4-46
Figure 4-6 The weighted response ratio for the responses to nitrogen
addition for fluxes and pools related to the ecosystem nitrogen
cycle in agricultural (open bars) and nonagricultural (closed
bars) ecosystems.	4-47
4.3.7.	Decomposition	4-47
Table 4-8 Decomposition.	4-49
4.3.8.	Nitrogen Mineralization	4-56
Table 4-9 Nitrogen mineralization.	4-58
Figure 4-7 Effects of nitrogen inputs on soil carbon, carbon:nitrogen ratio,
and minimum nitrogen in forest floors (panels A-C) and
mineral soils (panels D-F).	4-60
4.3.9.	Dissolved Organic Carbon Leaching	4-61
Table 4-10 Terrestrial dissolved organic carbon (DOC) leaching.	4-62
Figure 4-8 Conceptual diagram of positive (solid arrows) and negative
(dashed arrows) fluxes in nitrogen pools (squares) and carbon
pools (ovals) and the biological processes (no border) that are
affected by experimental nitrogen deposition.	4-67
4.3.10.	Belowground Carbon Pools 	4-67
Figure 4-9 Estimation of the changes in carbon budget of terrestrial
ecosystem under nitrogen addition.	4-69
Table 4-11 Belowground carbon pools.	4-70
4.3.11.	New Biogeochemical Indicators	4-72
Table 4-12 New biogeochemistry indicators.	4-73
4.3.12.	Differential Effects of Reduced and Oxidized nitrogen (N)	4-74
June 2018
x
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Table 4-13 The effects of different forms of inorganic nitrogen on
biogeochemical processes and indicators according to meta-
analyses. See Table 6-1 for the effects of different forms of
inorganic nitrogen on biological endpoints.	4-75
4.4.	Soil Monitoring and Databases	4-76
Table 4-14 Biogeochemistry monitoring and databases.	4-77
4.5.	Models	4-80
4.5.1.	Updates to Key Previously Identified Models	4-80
Table 4-15 Photosynthesis and Evapo Transpiration-Biogeochemical
(PnET-BGC) and DayCent.	4-87
4.5.2.	New Models (Published since 2008)	4-89
4.5.3.	Comparative Analyses	4-91
Table 4-16 Overview of properties of four dynamic soil chemistry models
as characterized in Bonten etal. (2015).	4-93
Table 4-17 Model comparison. 	4-95
4.6.	National-Scale Sensitivity	4-98
4.6.1.	Acidification Recovery	4-98
Table 4-18 Recovery. 	4-100
4.6.2.	Critical Loads	4-103
Figure 4-10 Forest ecosystem critical loads for soil acidity related to base
cation soil indicators.	4-104
Figure 4-11 Map of critical loads for nitrate leaching by ecoregion in the
U.S.	4-106
4.7.	Modification of Terrestrial Soil Response to Nitrogen (N)	4-106
4.7.1.	Disturbance and Stand Age Effects on Nitrogen Retention	4-107
4.7.2.	N and P Interactions	4-107
4.7.3.	Climate Modification of Acidification Effects on Soil	4-108
4.7.4.	Climate Modification of Nitrogen-Driven Eutrophication in Soil	4-109
Figure 4-12 The potential mechanisms that regulate the responses of
carbon dioxide (CO2), methane (CH4), and nitrous oxide (N2O)
production and consumption to elevated nitrogen (N). 	4-112
4.8.	Summary	4-112
4.8.1.	Sources	4-112
4.8.2.	Soil Processes and Indicators	4-113
Table 4-19 Summary of key soil geochemical processes and indicators
associated with eutrophication and acidification.	4-115
4.8.3.	Monitoring	4-118
4.8.4.	Models 	4-119
4.8.5.	National-Scale Sensitivity	4-121
4.8.6.	Climate Modification of Soil Response to N Addition 	4-122
APPENDIX 5. BIOLOGICAL EFFECTS OF TERRESTRIAL ACIDIFICATION 	5-1
5.1. Introduction	5-1
Figure 5-1 Diagram based on Fenn et al. (2006) showing indicators of
forest physiological function, growth, and structure that are
linked to biogeochemical cycles through processes that control
rates of calcium supply. Calcium affects plant physiological
processes that influence growth rates and the capacity of
plants to resist environmental stresses, such as extremes of
temperature, drought, insects, and diseases. Therefore,
acidifying deposition, which can deplete soil calcium or
interfere with calcium uptake through mobilization of soil
aluminum, can affect forest health.	5-3
June 2018
xi
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Table 5-1 Relationships between soil chemistry indicators and biological
endpoints that have been evaluated in the literature since the
2008 Integrated Science Assessment. 	5-4
5.2.	Effects on Terrestrial Organisms and Ecosystems	5-6
5.2.1.	Trees and Forests	5-6
Table 5-2 Summary of calcium addition studies in North America.	5-9
Figure 5-2 Relationship between the proportion of seedlings that were
sugar maple and soil base saturation in the upper B-horizon.	5-13
5.2.2.	Forest Understory and Grassland Species	5-20
5.2.3.	Lichens	5-21
5.2.4.	Soil Biota	5-22
5.2.5.	Fauna	5-24
5.3.	Characteristics, Distribution, and Extent of Sensitive Ecosystems	5-25
5.4.	Application of Terrestrial Acidification Models	5-26
5.5.	Levels of Deposition at Which Effects Are Manifested	5-27
5.5.1.	Impacts of Elevated Nitrogen and Sulfur Deposition	5-27
Table 5-3 Impacts of acidifying nitrogen and sulfur deposition.	5-30
5.5.2.	Impacts of Ambient Deposition	5-36
Table 5-4 Results of Spearman's rank correlation analysis comparing
growth versus critical load exceedance by species for the
forested plots in the northeastern U.S. [from Duarte et al.
(2013)]. Modeled sulfur and nitrogen deposition on the plots
ranged from 242 to 1,154 eq/ha/yr and 256 to 920 eq/ha/yr,
respectively. Correlations shown here are significant at
a = 0.05. 	5-37
Table 5-5 Results of Spearman's rank correlation analyses comparing
tree vigor, crown density and dieback, and canopy
transparency versus critical load exceedance by species for
the forested plots in the northeastern U.S. [from Duarte et al.
(2013)]. Modeled sulfur and nitrogen deposition on the plots
ranged from 242 to 1,154 eq/ha/yr and 256 to 920 eq/ha/yr,
respectively. Correlations shown here are significant at
a = 0.05. 	5-38
5.5.3.	Critical Loads and Exceedances	5-40
5.6.	Climate Modification of Ecosystem Response	5-46
5.7.	Summary	5-48
5.7.1.	Physiology and Growth	5-48
5.7.2.	Biodiversity	5-50
5.7.3.	National-Scale Sensitivity and Critical Loads	5-50
Table 5-6 Mode of action for acidifying nitrogen and sulfur deposition.	5-52
APPENDIX 6. TERRESTRIAL ECOSYSTEMS: NITROGEN ENRICHMENT EFFECTS
ON ECOLOGICAL PROCESSES	6-1
6.1.	Introduction	6-1
6.2.	Linking Nitrogen Deposition to Changes in Physiology, Growth, and Productivity in
Terrestrial Ecosystems	6-2
6.2.1.	Introduction	6-2
6.2.2.	Mechanisms Operating across Terrestrial Ecosystems 	6-3
Figure 6-1 Effects of nitrogen additions on plant growth and net primary
productivity.	6-8
Figure 6-2 Effects of added nitrogen on ecosystem carbon pools and
fluxes. 	6-9
Table 6-1 The effects of different forms of inorganic nitrogen on biological
endpoints according to meta-analyses. See Table 4-13 for the
effects of different forms of inorganic nitrogen on
biogeochemical processes and indicators. 	6-14
June 2018	xii	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
6.2.3. Forests
6-17
Table 6-2
Table 6-3
Table 6-4
Table 6-5
Figure 6-3
Growth, productivity, and carbon cycle responses of
ectomycorrhizal fungi to nitrogen added via atmospheric
deposition or experimental treatments.	
Growth, productivity, and carbon cycle responses of arbuscular
mycorrhizal fungi to nitrogen added via atmospheric deposition
or experimental treatments.	
Abundance and carbon cycle responses of forest soil
microorganisms and soil invertebrates to nitrogen added in
experimental treatments.	
Growth and physiology responses of forest epiphytic lichens to
nitrogen added via atmospheric deposition or experimental
treatments.	
Studies reporting the response of forest (A) aboveground
biomass carbon sequestration and (B) ecosystem carbon
sequestration to nitrogen deposition or long-term nitrogen
additions.
6.2.4. Arctic and Alpine Tundra and Grasslands_
6.2.5.
6.2.6.
Table 6-6
Table 6-7
Table 6-8
Table 6-9
Grasslands
Table 6-10
Alpine and Arctic tundra plant productivity and physiology
responses to nitrogen added via atmospheric deposition or
experimental treatments.	
Alpine and Arctic tundra lichen growth and physiology
responses to nitrogen added via atmospheric deposition or
experimental treatments.	
Growth and biodiversity responses of ericoid mycorrhizal fungi
to nitrogen added in experimental treatments.	
Alpine and Arctic tundra microbial biomass responses to
experimental nitrogen additions.	
Grassland microbial biomass responses to experimental
nitrogen additions.	
Arid and Semiarid Ecosystems	
Table 6-11 Arid and semiarid ecosystem plant productivity and physiology
Table 6-12
responses to nitrogen added in experimental treatments.
Arid and semiarid microbial biomass responses to
experimental nitrogen additions.	
6.3. Relationships between Nitrogen Deposition and Terrestrial Species Composition,
Species Richness, and Biodiversity	
6.3.1.
6.3.2.
6.3.3.
Introduction	
Mechanisms Operating across Terrestrial Ecosystems
Forests	
Table 6-13
Forest plant diversity responses to nitrogen added via
atmospheric deposition or experimental treatments.	
Table 6-14
Table 6-15
Table 6-16
Forest microbial biodiversity responses to nitrogen added via
atmospheric deposition or experimental treatments.	
Ectomycorrhizal biodiversity responses to nitrogen added via
atmospheric deposition or experimental N additions. 	
Arbuscular mycorrhizal responses to nitrogen added via
atmospheric deposition or experimental treatments.	
6.3.4.
Table 6-17 Arthropod and other invertebrate responses to experimental
nitrogen additions.	
Alpine and Arctic Tundra	
Table 6-18 Alpine and Arctic tundra plant diversity responses to nitrogen
6-20
6-30
6-32
6-39
.6-46
6-46
6-49
.6-67
.6-71
.6-72
.6-74
.6-80
.6-82
.6-85
6-95
_ 6-97
_ 6-97
_ 6-98
6-102
6-105
6-112
6-116
6-121
6-125
6-129
Table 6-19
6.3.5. Grasslands
added via atmospheric deposition or experimental treatments.	6-131
Alpine and Arctic tundra microbial diversity responses to
nitrogen added via experimental treatments.	6-136
6-138
June 2018
xiii
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Table 6-20 Grassland microbial diversity responses to nitrogen added via
experimental treatments.	6-145
6.3.6.	Arid and Semiarid Ecosystems	6-147
Table 6-21 Arid and semiarid ecosystem plant diversity responses to
nitrogen added via atmospheric deposition or experimental
treatments.	6-149
Table 6-22 Arid and semiarid ecosystem microbial diversity responses to
nitrogen added via experimental treatments.	6-156
6.3.7.	Lichens	6-157
Table 6-23 Lichen biodiversity responses to nitrogen added via
atmospheric deposition or experimental treatments.	6-158
6.3.8.	Most Sensitive Ecosystems 	6-161
6.4.	Climate Modification of Ecosystem Nitrogen Response	6-162
6.5.	Critical Loads	6-164
Figure 6-4 Comparison of European and U.S. empirical critical loads for
nitrogen from Pardo et al. (2011a).	6-167
6.5.1.	Mycorrhizal Fungi	6-167
Figure 6-5 Map of critical loads for mycorrhizal fungi by ecoregion in the
U.S.	6-169
Table 6-24 Mycorrhizal critical loads.	6-170
6.5.2.	Lichens and Bryophytes	6-170
Figure 6-6 Map of critical loads for lichens by ecoregion in the U.S.	6-172
Table 6-25 Lichen critical loads. 	6-173
6.5.3.	Herbaceous and Shrub Species	6-175
Table 6-26 Herbaceous and shrub critical loads.	6-177
Figure 6-7 Map of critical loads for herbaceous plants and shrubs by
ecoregion in the U.S.	6-180
Figure 6-8 Nitrogen deposition (gray scale) and critical loads for nitrogen
deposition based on total graminoid plus forb species richness
(colored symbols).	6-181
6.5.4.	Trees	6-183
Table 6-27 Tree critical loads.	6-184
Figure 6-9 Map of critical loads for forest ecosystems by ecoregion in the
U.S.	6-185
6.5.5.	Critical Loads Exceedance Studies	6-186
6.6.	Summary	6-187
6.6.1.	Physiology, Growth, and Productivity Summary	6-187
6.6.2.	Biodiversity Summary	6-190
6.6.3.	Critical Loads Summary	6-192
Table 6-28 Critical loads for nitrogen by Pardo et al. (2011c) and more
recent critical load information. The critical loads (CLs)
reported for Pardo et al. (2011c) are for Level 1 Ecoregions,
whereas the new CL information may be for more specific
ecosystems within each respective Ecoregion (e.g., California
coastal sage scrub within the Mediterranean California
Ecoregion; this is noted in the 5th column of the table entitled
"Type of Ecosystem").	6-194
APPENDIX 7. AQUATIC BIOGEOCHEMISTRY	7-1
7.1. Biogeochemistry of Nitrogen and Sulfur in Freshwater Systems	7-1
Table 7-1 Summary of key freshwater indicators of eutrophication and
acidification.	7-4
7.1.1. Nitrogen and Sulfur Sources	7-4
Table 7-2 Summary of recent studies quantifying nitrogen deposition
contribution to total nitrogen loading in freshwater systems.	7-7
June 2018	xiv	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
7.1.2.	Ecosystem Processes, Effects, and Indicators	7-11
Figure 7-1 Nitrogen cycle in freshwater ecosystem.	7-12
Figure 7-2 Quarterly measured concentrations of a range of water
chemistry variables at Blue Lough from 1990 to 2014.	7-17
Figure 7-3 Dissolved organic carbon (DOC) concentrations from 1993 to
2013 for Bracey Pond (a), Salmon Pond (b), Jellison Pond (c),
Second Pond (d), Little Long Pond (e), and Tilden Pond (f) in
Maine.	7-31
7.1.3.	Freshwater Monitoring and Databases	7-34
Table 7-3 Monitoring and resurvey results of aquatic acidification and/or
chemical recovery since the 2008 Integrated Science
Assessment for Oxides of Nitrogen and Sulfur—Ecological
Criteria.	7-36
7.1.4.	Models 	7-43
Table 7-4 Recent process-based model estimates of surface water
acidification and chemical recovery in the U.S.	7-45
7.1.5.	National Scale Sensitivity and Response	7-46
Table 7-5 Model projections of surface water sulfate and associated acid
neutralizing capacity, shown as changes between dates, for
Adirondack and Shenandoah streams.	7-48
7.1.6.	Water Quality Criteria	7-58
Table 7-6 Numeric nutrient water quality criteria for rivers/streams by
state (all values in mg/L).	7-60
Table 7-7 U.S. EPA aggregate Level III ecoregion nutrient criteria (all
values in mg/L; U.S. EPA ecoregional nutrient criteria
documents for rivers and streams).	7-64
Figure 7-4 Total nitrogen criterion values by ecoregion.	7-65
Figure 7-5 Chlorophyll a criterion values by ecoregion. 	7-65
7.1.7.	Freshwater Biogeochemistry Summary	7-66
7.2. Biogeochemistry of Nitrogen in Estuarine and Near-Coastal Systems	7-69
Table 7-8 Summary of key indicators of nitrogen enrichment and
nutrient-enhanced coastal acidification.	7-70
7.2.1.	Nitrogen Sources 	7-71
Figure 7-6 Chemical nitrogen cascade in the Chesapeake Bay watershed
(metric tons/year).	7-73
Table 7-9 Summary of studies quantifying atmospheric nitrogen
contribution to total nitrogen in coastal areas via watersheds
and/or direct deposition to estuary surface waters.	7-74
7.2.2.	The Estuary Environment	7-78
Figure 7-7 Schematic diagram illustrating sources, transformations, and
fate of nitrogen along the estuary-to-ocean continuum.
Surface, subsurface, and atmospheric pathways of externally
supplied or new nitrogen inputs attributable to anthropogenic
activities are shown as internal nitrogen cycling. The combined
anthropogenic nitrogen inputs are shown as a thick arrow
(upstream), which decreases in thickness downstream as a
portion of the nitrogen inputs settles to the bottom sediments
and is buried and/or denitrified. Nitrogen (N2) fixation is a
biologically mediated new nitrogen input. The linkage of
anthropogenically enhanced nitrogen inputs to accelerated
primary production or eutrophication and its trophic and
biogeochemical fate are also shown.	7-79
7.2.3.	Dissolved Oxygen and Hypoxia	7-79
7.2.4.	Estuarine and Near-Coastal pH 	7-80
7.2.5.	Nitrogen in Surface Waters	7-81
7.2.6.	Nitrogen Cycling	7-82
June 2018
xv
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 7-8 New complexities in nitrogen cycling have been detailed since
the 2008 ISA as shown in this illustration of the sedimentary N
cycle in the Lower St. Lawrence estuary. 	7-83
7.2.7.	Monitoring Data	7-94
Figure 7-9 Percentage of area in each coastal region scoring good, fair,
and poor based on the Water Quality Index for the NCCA
2010.	7-95
7.2.8.	Modeling Estuaries and Near-Coastal Areas	7-98
7.2.9.	National Scale Sensitivity	7-102
7.2.10.	Water Quality Criteria for Estuaries	7-104
Figure 7-10 State progress toward developing numeric nutrient criteria for
estuaries https://www.epa.gov/nutrient-policy-data/state-
progress-toward-developing-numeric-nutrient-water-quality.	7-105
7.2.11.	Estuary and Near-Coastal Biogeochemistry Summary	7-106
APPENDIX 8. BIOLOGICAL EFFECTS OF FRESHWATER ACIDIFICATION 	8-1
8.1.	Introduction	8-1
8.2.	Chronic versus Episodic Acidification	8-3
8.3.	Aquatic Organisms Impacted by Acidifying Deposition	8-5
8.3.1.	Plankton	8-5
8.3.2.	Periphyton	8-9
8.3.3.	Benthic Invertebrates	8-10
Figure 8-1 Total macroinvertebrate species (community) richness as a
function of median pH in 36 streams sampled in the western
Adirondack Mountains of New York, 2003-2005; the four
standard (New York State) impact categories for species
richness are defined.	8-12
Table 8-1 Thresholds of biological response to changes in water acidity
for benthic invertebrates published since the 2008 Integrated
Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria. 	8-13
8.3.4.	Bacteria, Macrophytes, and Bryophytes	8-14
8.3.5.	Amphibians	8-15
8.3.6.	Fish	8-16
Figure 8-2 Relationship between (a) inorganic monomeric aluminum and
gill aluminum for parr and smolt, and (b) acid neutralizing
capacity and gill aluminum.	8-20
Figure 8-3 Critical aquatic pH ranges for fish species.	8-22
Table 8-2 pH thresholds in fish published since the 2008 Integrated
Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria. 	8-24
Figure 8-4 Number of fish species per lake verses acidity status,
expressed as acid neutralizing capacity, for Adirondack lakes.	8-25
Table 8-3 Expected ecological effects and concern levels in freshwater
ecosystems at various levels of acid neutralizing capacity.	8-27
Table 8-4 Threshold values of aluminum for various fish species and
associated effects.	8-29
Figure 8-5 Relationship between (a) pH, (b) cationic aluminum, (c) acid
neutralizing capacity, and (d) gill aluminum as compared with
accumulated mortality of Atlantic salmon smolt.	8-33
8.3.7.	Fish-Eating Birds	8-34
8.3.8.	Aquatic Assemblages	8-34
Figure 8-6 Species richness of biotic groups in 30 Adirondack study lakes
relative to midsummer epilimnetic pH during sample years.	8-36
June 2018	xvi	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 8-7 Structure of diatom assemblage in 20 streams across a pH
gradient of 4.5-8.5. (a) Species richness (log-io-transformed
numbers of diatom species per stream), (b) total abundance of
diatoms (log-io-transformed numbers of individuals per m2),
(c) chlorophyll a concentration (mg chlorophyll a per m2), as a
measure of algal biomass.	8-37
Figure 8-8 Macroinvertebrate community composition in 20 streams
across a pH gradient. Taxon richness (total number of primary
consumer taxa; left panel) and benthic density [number of
individuals logio(x+ 1)-transformed; right panel] plotted against
stream pH, all primary consumers (a, b), shredders (excluding
Leuctridae and Nemouridae) (c, d), herbivore-detritivores
(Leuctridae and Nemouridae) (e, f), collectors (g, h) and
grazers (i, j).	8-38
8.4.	Documentation of Biological Recovery	8-39
8.4.1.	Phytoplankton Recovery	8-42
Table 8-5 Paleolimnological responses to changing lake chemistry
published since the 2008 Integrated Science Assessment for
Oxides of Nitrogen and Sulfur-Ecological Criteria.	8-43
8.4.2.	Zooplankton Recovery	8-44
Figure 8-9 (A) Midsummer sulfate concentration in the epilimnion of
Brooktrout Lake (o) and in annual wet deposition (•) at local
National Atmospheric Deposition Program/National Trends
Network Station NY52 from 1984-2012. (B) Midsummer
epilimnetic concentrations of inorganic monomeric aluminum
(~) and [hydrogen ion] (¦) in Brooktrout Lake from 1984-2012.
(C) Midsummer phytoplankton (A) and total plankton
(phytoplankton, rotifers, crustaceans) (A) species richness in
Brooktrout Lake from 1984-2012.	8-46
Table 8-6 Midsummer values and productivity analytes in the epilimnion
of Brooktrout Lake in the Adirondack Park from the 1980s
through 2010-2012.3	8-47
8.4.3.	Benthic Invertebrate Recovery	8-50
8.4.4.	Fish and Amphibian Recovery	8-53
8.4.5.	Bird Recovery	8-55
8.4.6.	Mitigation	8-56
8.5.	Levels of Deposition at Which Effects Are Manifested	8-58
8.5.1.	Most Sensitive and Most Affected Ecosystems and Regions	8-58
Figure 8-10 Map of landscape sensitivity to acidic deposition for the
northeastern and mid-Atlantic U.S. Stippled areas were not
considered.	8-59
8.5.2.	Extent and Distribution of Sensitive Ecosystems/Regions	8-60
Figure 8-11. Surface water Acid Neutralizing Capacity (ANC) map, based on
data compiled by Sullivan (2017)	8-62
Figure 8-12 (a) Minimum critical loads of surface water acidity for nitrogen
and sulfur. Grids represent the minimum calculated critical load
from all data within the 36 * 36-km grid cell (b) Mean critical
loads of surface water acidity. Grids represent the average
calculated critical load from all data within the 36 * 36-km grid
cell. The critical chemical criterion used was an acid
neutralizing capacity 50 peq/L. 	8-63
8.5.3.	Climate Modification of Ecosystem Response to N and S 	8-64
8.5.4.	Critical Loads	8-64
Table 8-7 Recent empirical critical loads to protect against aquatic
acidification in U.S. ecosystems.	8-67
Table 8-8 Recent aquatic critical load and target load modeling studies in
the U.S. to protect against aquatic acidification.	8-68
June 2018	xvii	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 8-13 Target loads for sulfur deposition in the Adirondack Park to
protect lake acid neutralizing capacity to 50 peq/L in the
Year 2010 (left map) and their exceedance (right map).	8-74
Figure 8-14. Modeled target loads and median measured Acid Neutralizing
Capacity (ANC) in the period 1993 to 1996 from 30 Great
Smoky Mountain National Park (GRSM) streams to achieve an
ANC of 20 peq/L by the Year 2150.	8-77
Figure 8-15. Exceedance level of current NO3" + SO42" atmospheric
deposition for 387 stream sites in the GRSM. Exceedances
were calculated for the years 2050 and 2150 using two targets
for modeled stream ANC recovery of 20 pmolc/L and
20 pmolc/L less than the simulated preindustrial ANC.	8-78
8.6. Aquatic Acidification Summary and Causal Determinations	8-80
Table 8-9 Ecological indicators for aquatic acidification.	8-82
8.6.1.	Phytoplankton	8-82
8.6.2.	Zooplankton 	8-83
8.6.3.	Benthic Invertebrates	8-83
8.6.4.	Fish	8-84
8.6.5.	Thresholds of Response	8-85
Table 8-10 Results of recent biological effects studies in surface waters
indicative of thresholds of biological response to changes in
water acidity.	8-86
8.6.6.	Biological Recovery	8-86
8.6.7.	Most Sensitive and Most Affected Regions	8-87
8.6.8.	Critical Loads	8-88
APPENDIX 9. BIOLOGICAL EFFECTS OF FRESHWATER NITROGEN ENRICHMENT	9-1
9.1.	Introduction to Nitrogen Enrichment and Eutrophication in Freshwater Systems	9-1
Figure 9-1 Conceptual model of the influence of atmospheric nitrogen
deposition on freshwater nutrient enrichment.	9-2
9.1.1. Deposition to Freshwater Systems	9-5
Table 9-1 Summary of studies using diatoms as biological indicators of
nitrogen enrichment evaluated in the literature since the 2008
Integrated Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria. 	9-9
Figure 9-2 Phytoplankton responses to nitrogen and/or phosphorus
enrichment for Rocky Mountain lakes receiving low or high
atmospheric nitrogen deposition, given as the ratio of final
chlorophyll concentration in the enriched treatment
(+phosphorus, +nitrogen, or+nitrogen + phosphorus) to the
chlorophyll concentration in the unenriched control.	9-12
9.2.	Biological Indicators	9-15
9.2.1.	Diatoms	9-16
9.2.2.	Ratios of Nitrogen and Phosphorus 	9-17
9.2.3.	Phytoplankton Biomass Nitrogen (N) to Phosphorus (P) Limitation Shift	9-18
9.2.4.	Periphyton/Microbial Biomass	9-18
9.2.5.	Chlorophyll a	9-20
Table 9-2 Summary of additional evidence for nitrogen limitation on
productivity of freshwater ecosystems that has been evaluated
since the 2008 Integrated Science Assessment for Oxides of
Nitrogen and Sulfur-Ecological Criteria.	9-24
9.2.6.	Potential Biological Indicators	9-27
9.3.	Community Composition, Species Richness, and Diversity	9-29
9.3.1.	Archaea and Bacterial Diversity	9-30
9.3.2.	Phytoplankton Diversity	9-30
June 2018
xviii
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Table 9-3 Summary of studies on nitrogen effects on species
composition and biodiversity that have been evaluated in the
literature since the 2008 Integrated Science Assessment for
Oxides of Nitrogen and Sulfur-Ecological Criteria.	9-31
9.3.3.	Benthic Algal Diversity	9-39
9.3.4.	Zooplankton Diversity	9-40
9.3.5.	Macroinvertebrate Diversity 	9-41
9.3.6.	Macrophyte Diversity	9-42
9.3.7.	Amphibian Diversity	9-43
9.3.8.	Fish Diversity	9-43
9.4.	Animal Behavior and Disease	9-43
9.4.1.	Behavior	9-43
9.4.2.	Disease	9-44
9.5.	Summary of Thresholds, Levels of Deposition at Which Effects Are Manifested, and
Critical Loads	9-44
Table 9-4 Summary of critical loads for nitrogen eutrophication for
surface water in the U.S. [adapted from Pardo et al. (2011c)
with newer studies added], 	9-46
Table 9-5 Summary of mean lake nitrate (NO3") concentrations, inorganic
nitrogen deposition amounts, and nutrient enrichment inflection
points where lake NO3" concentrations reflect increased
nitrogen deposition, [from Baron etal. (2011b)].	9-49
9.6.	Summary and Causal Determination	9-49
APPENDIX 10. BIOLOGICAL EFFECTS OF NITROGEN ENRICHMENT IN ESTUARIES
AND NEAR-COASTAL SYSTEMS	10-1
10.1.	Introduction	10-1
Figure 10-1 Eutrophication can occur when the availability of nutrients
increases above normal levels.	10-3
Figure 10-2 Overall eutrophication condition on a national scale.	10-5
10.1.1.	Nitrogen Sources to Estuaries and Coasts	10-5
10.1.2.	Trends in Atmospheric Deposition of Nitrogen (N)	10-6
10.1.3.	Nitrogen Limitation	10-6
10.1.4.	Characteristics of Coastal Systems Sensitive to Eutrophication	10-8
10.2.	Indicators of Nutrient Enrichment	10-11
Table 10-1 Indicators of Estuarine Eutrophication. 	10-12
Figure 10-3 Biological indicator responses to nutrient enrichment.	10-13
10.2.1.	Chlorophyll a	10-13
Table 10-2 Chlorophyll a thresholds used in methods to evaluate the
status of phytoplankton in U.S. coastal and estuarine water
bodies.	10-14
10.2.2.	Harmful/Nuisance/Toxic Algal Blooms	10-16
Table 10-3 Levels and forms of nitrogen at which effects on phytoplankton
are manifest in U.S. coastal waters evaluated in the literature
since the 2008 Integrated Science Assessment for Oxides of
Nitrogen and Sulfur-Ecological Criteria.	10-20
10.2.3.	Macroalgal Abundance	10-25
10.2.4.	Dissolved Oxygen	10-26
Figure 10-4 The range of ecological impacts exhibited as dissolved oxygen
levels drop from saturation to anoxia.	10-27
Figure 10-5 Coastal eutrophic and hypoxic areas of North America the
Caribbean.	10-29
10.2.5.	Submerged Aquatic Vegetation 	10-32
Table 10-4 Nitrogen loading thresholds from multiple watershed sources
versus eelgrass loss.	10-34
June 2018
xix
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 10-6 Extent of submerged aquatic vegetation in the Chesapeake
Bay 1978-2016. 	10-36
10.2.6. Indices of Estuarine Condition 	10-37
10.3.	Effects on Biodiversity	10-39
10.3.1.	Paleontological Diversity	10-39
10.3.2.	Phytoplankton Diversity	10-40
10.3.3.	Diversity of Phytoplankton Under Different Forms of Nitrogen (Reduced vs.
Oxidized)	10-41
Figure 10-7 Summary conceptual schematic illustrating the effect of
changes in the proportion of NhV and NO3" in the loads of N
provided to a natural system. When NhVis the dominant form,
and when waters are warmer, flagellates, cyanobacteria, and
chlorophytes among other classes may proliferate, leading to
overall productivity dominated by the small size class of algae
(e.g., <5 |jm). In contrast, when NO3" is the dominant form
provided, especially under cooler water conditions, diatoms
more likely dominate, and their overall production will be more
likely dominated by cells of a larger size class (e.g., >5 |jm).
Moreover, chlorophyll a yield and total production may be
higher than under the NH4+enrichment condition.	10-43
10.3.4.	Diversity of Bacteria, Archaea, and Microzooplankton	10-44
10.3.5.	Benthic Diversity	10-45
10.3.6.	Fish Diversity	10-46
10.3.7.	Trophic Interactions	10-46
10.3.8.	Models Linking Indicators to Nitrogen Enrichment	10-47
Figure 10-8 Forecasting curves for effects on total nitrogen loadings on
(A) chlorophyll and (B) dissolved oxygen (mean and 95%
confidence interval) for selected estuaries demonstrating the
full range of sensitivity to relative total nitrogen loading.	10-49
10.4.	Animal Behavior and Disease	10-50
10.4.1.	Behavior	10-50
Figure 10-9 The pathway of effects of eutrophication on different
reproductive behaviors and selection forces in Gasterosteus
aculeatus. 	10-51
10.4.2.	Disease	10-52
10.5.	Nutrient Enhanced Coastal Acidification	10-53
Figure 10-10 Pathway of nutrient-enhanced coastal acidification from
nitrogen loading to biological effects. Both microbial respiration
of organic matter and increasing atmospheric carbon dioxide
lower pH of coastal waters.	10-55
10.6.	Summary of Thresholds and Levels of Deposition at Which Effects Are Manifested	10-58
10.7.	Summary and Causal Determinations	10-59
10.7.1.	Nitrogen Enrichment	10-60
10.7.2.	Nutrient-Enhanced Coastal Acidification	10-63
APPENDIX 11. NITROGEN ENRICHMENT EFFECTS IN WETLANDS	11-1
11.1.	Introduction	11-1
Table 11-1 Wetland classification used in the Integrated Science
Assessment.	11-2
11.2.	Regional Sensitivity 	11-4
Figure 11-1 Estimated extent of wetlands stressed by nonnative plants as
determined by Nonnative Plant Stressor Indicator, at a national
or regional level. 	11-5
11.2.1. Climate Modification of Ecosystem Response to Nitrogen	11-6
11.3.	Soil Biogeochemistry	11-6
11.3.1. Nitrogen Pools and Processes	11-7
June 2018	xx	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Table 11-2 New studies on nitrogen addition effects on nitrogen cycling in
wetlands.	11-13
11.3.2. Soil Carbon Cycling	11-19
Table 11-3 Loading effects upon belowground carbon cycling.	11-25
Table 11-4 Nitrogen loading effects upon methane emissions.	11-34
11.4.	Production and Aboveground Biomass	11-36
11.4.1.	Salt Marsh	11-36
11.4.2.	Mangrove	11-38
11.4.3.	Freshwater Tidal Marsh 	11-38
11.4.4.	Intermittent Wetland	11-39
11.4.5.	Bog and Fen	11-39
11.4.6.	SummaryTable	11-41
Table 11-5 Nitrogen loading effects upon production and biomass. 	11-41
11.5.	Plant Stoichiometry and Physiology	11-48
11.5.1.	Salt Marsh	11-49
11.5.2.	Mangrove	11-49
11.5.3.	Freshwater Marsh	11-50
11.5.4.	Riparian Wetland	11-51
11.5.5.	Bog and Fen	11-51
11.5.6.	SummaryTable	11-56
Table 11-6 Nitrogen loading effects upon plant stoichiometry and
physiology.	11-56
11.6.	Plant Architecture	11-68
11.6.1.	Salt Marsh	11-68
11.6.2.	Mangrove	11-69
11.6.3.	Freshwater Tidal Marsh 	11-69
11.6.4.	Riparian Wetland	11-70
11.6.5.	SummaryTable	11-70
Table 11-7 Nitrogen loading effects upon architecture.	11-70
11.7.	Demography	11-73
11.7.1.	Mangrove	11-73
11.7.2.	Riparian Wetland	11-73
11.7.3.	SummaryTable	11-74
Table 11-8 Nitrogen loading effects upon demography. 	11-74
11.8.	Biodiversity/Community	11-74
11.8.1.	Plants	11-75
Table 11-9 Nitrogen loading effects upon plant biodiversity and
communities.	11-79
11.8.2.	Consumers	11-83
11.9.	Critical Loads	11-85
11.9.1.	Freshwater Wetland	11-85
11.9.2.	Coastal Wetlands	11-85
11.9.3.	Comparison to Critical Loads from Europe	11-86
11.10.	Summary	11-87
11.10.1. Causality across Wetland Types	11-87
Figure 11-2 Summary of the levels of nitrogen addition where a change to
nitrogen cycling is observed.	11-89
Figure 11-3 Summary of new literature of nitrogen load effects on
belowground and aboveground carbon cycling. 	11-91
Figure 11-4 Summary of the level of nitrogen load that caused a change in
the response variables of plant stoichiometry and physiology in
wetlands.	11-93
Figure 11-5 Summary of nitrogen addition studies on wetland biodiversity.	11-95
June 2018
xxi
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
11.10.2. Coastal versus Freshwater Wetlands	11-96
Figure 11-6 Summary of field nitrogen addition studies for coastal wetlands
(blue borders) versus critical load (black border).	11-97
Figure 11-7 Summary of nitrogen load studies for freshwater wetlands as
well as current critical loads.	11-98
APPENDIX 12. NONACID SULFUR ENRICHMENT EFFECTS	12-1
12.1.	Introduction	12-2
Figure 12-1 Effects of sulfur oxide deposition on chemical (blue boxes),
biological (green boxes), and atmospheric (yellow boxes)
indicators of ecosystem change, as documented by the
previous Integrated Science Assessment and more recent
research.	12-3
12.2.	Ecosystem Effects of Altered Sulfur Cycling	12-4
12.2.1.	The Sulfur Cycle	12-4
Figure 12-2 Sulfur cycle in terrestrial, forested ecosystems.	12-4
12.2.2.	Deposition and Sulfur Stores 	12-8
Table 12-1 New study on sulfur (S) deposition effects on sulfur (S) cycling.	12-9
12.2.3.	Sulfide Toxicity	12-9
Table 12-2 Quantitative effects of sulfide on wetland and aquatic plant
species.	12-11
Figure 12-3 Schematic from Minnesota Pollution Control Agency that
illustrates the mitigating effect of iron on the toxicity of sulfide
and the stimulatory effect that organic carbon has on sulfide
production.	12-14
12.2.4.	Internal Eutrophication	12-15
Figure 12-4 Mechanisms of linked sulfur, iron, and phosphorus cycling in
wetland waters and soils.	12-17
12.2.5.	Sulfur Effects on Methane Emissions	12-18
Table 12-3 New studies on nonacidifying sulfur effects on methane
emissions.	12-20
12.3.	Interactions between Sulfur Deposition and Mercury	12-22
Figure 12-5. Sulfate, dissolved organic carbon, and inorganic mercury are
all important determinants of the rate and amount of methyl
mercury produced by sulfate-reducing prokaryotes.	12-24
12.3.1.	Mercury Cycle and the Importance of Methylation	12-24
Figure 12-6 Cycling of mercury (Hg) and methylmercury (MeHg) in
ecosystems.	12-25
12.3.2.	Biology of Sulfate-Reducing Prokaryotes	12-27
Table 12-4 New studies on the biology of sulfate-reducing prokaryotes.	12-28
Figure 12-7 The effect of different environmental factors on the relationship
between sulfate and mercury methylation. Methylmercury
(MeHg) accumulation is minimal at low and high sulfate
concentrations, with an optimum near 100 |jM sulfate (blue
line). High dissolved organic matter (DOM) will increase the
magnitude of MeHg production across the range of sulfate
concentrations (red line). Sulfide produced by methylators will
inhibit further methylation if it accumulates in the aqueous
environment where methylation occurs, shifting the MeHg
optimum left (purple line). However if ecosystem chemistry
(iron [Fe], reoxidation of sulfide, organic matter [OM]) allows for
rapid sequestration of sulfide to large particles and sediments,
the relaxation of the negative feedback of sulfide to
sulfur-reducing prokaryotes (SRPs) will shift the MeHg
optimum right (green line).	12-36
12.3.3.	Environmental Drivers of Mercury Methylation Potential	12-37
Table 12-5 Environmental factors that affect mercury (Hg) methylation.	12-37
June 2018	xxii	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)

Figure 12-8 The relationship between surface water sulfate and total


mercury or methylmercury fraction in river/leaf litter


mesocosms.
12-41

Table 12-6 New mesocosm or incubation studies on sulfur addition effects


on methylmercury.
12-42

Figure 12-9 Total methylmercury mass in water at Little Rock Lake, Wl,


annually (a), and in relationship to annual mercury (b) or


sulfur (c) deposition.
12-47
12.3.4.
Mercury Methylation in Sulfur Addition Field Studies
12-54

Table 12-7 New studies on mercury methylation in sulfur-amended


ecosystems.
12-55

Table 12-8 New studies on mercury (Hg) and sulfur (S) cycling in the


Everqlades Water Conservation Areas (WCA).
12-60

Figure 12-10 The relationship between surface water sulfate and


methylmercury concentrations in the Florida Everqlades.
12-64
12.3.5.
Drivers of Mercury Methylation under Ambient Conditions
12-67

Table 12-9 New studies on correlations between sulfur (S) and


methylmercury (MeHq) in ecosystems.
12-69

Figure 12-11 Concentrations of soil sulfur (S), elevated by sulfur oxides


(SOx) deposition, correlate with rice grain (methylmercury


[MeHg]) and rice grain %MeHg in rice paddies near industrial


emitters in China.
12-75

Fiqure 12-12 Interactions between mercury (Hq) and sulfur (S) cycles.
12-75

Table 12-10 New studies on deposition of sulfur (S) and mercury (Hg) and


their effect on methylmercury (MeHq).
12-76

Table 12-11 New studies of Interactions between organic matter and sulfur


(S) in Controllinq methylmercury (MeHq).
12-79
Sulfur Impacts on Mercury in Wildlife
12-81

Figure 12-13 Bioconcentration and biomagnification result in methylmercury


concentrations about 1 million times higher in predator fish


than in stream water.
12-82

Table 12-12 New studies on sulfur (S) impacts upon mercury (Hg) in


wildlife.
12-83

Figure 12-14 Tissue mercury concentrations as a function of surface water


sulfate concentrations (n = 2,360 surface water samples) in the


Everqlades Protection Area.
12-88
Extent and Distribution of Sensitive Ecosystems
12-88

Fiqure 12-15 Fish mercury concentrations across the U.S.
12-90
Critical Loads
12-90
Summary of Nonacidifyinq Sulfur Effects
12-90

Table 12-13 Summary of quantitative effects of nonacidifying sulfur


enrichment.
12-91
12.7.1.
Terrestrial Sulfur Cyclinq
12-93
12.7.2.
Aquatic Sulfur Cyclinq
12-94
12.7.3.
Sulfide Toxicity
12-94

Figure 12-16 Thresholds of sulfate or sulfide concentrations in water, which


cause bioloqical and chemical effects in ecosystems.
12-96
12.7.4.
Internal Eutrophication
12-96
12.7.5.
Effects on Methane Production
12-96
12.7.6.
The Role of Microbes in Mercury Methylation
12-96

Figure 12-17 Linear relationships between sulfate and methylmercury


concentrations in published studies.
12-98
12.7.7.
Impacts of Sulfur upon Mercury Cyclinq
12-99

Figure 12-18 Thresholds of sulfate addition or deposition from published


studies which affect chemical or biological changes in


ecosystems.
12-100
June 2018
xxiii
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
12.7.8.	Sensitive Ecosystems	12-100
12.7.9.	Mercury Effects on Animal Species	12-101
Supplemental Figure 1 Mercury and methylmercury mass balance cycle in
Lake Onondaga in 1992. The quantities of mercury in each flux
are indicated in parentheses by (kg total Hg/yr, kg MeHg/yr).	12-107
APPENDIX 13. CLIMATE MODIFICATION OF ECOSYSTEM RESPONSE TO NITROGEN
AND SULFUR	13-1
13.1.	Climate Modification of Soil Acidification and Nitrogen Enrichment	13-2
13.1.1.	Nitrogen Transport and Transformation	13-2
Figure 13-1 Summary of key interactions between nitrogen,
anthropogenic-driven climate change, and hydrology.	13-3
Table 13-1 Summary on climate modification sulfur and nitrogen cycling in
Appendix 4 and Appendix 6 in addition to those in
Appendix 13.	13-6
13.1.2.	Nitrogen, Climate, and Carbon Cycling	13-7
Figure 13-2 The effects of increased nitrogen, temperature, and
precipitation on terrestrial carbon pools (left panel) and fluxes
(right panel) from published meta-analyses.	13-9
Table 13-2 Summary on climate modification of nitrogen (N) effects on
carbon (C) cycling in Appendix 4 and Appendix 6 in addition to
those in Appendix 13. 	13-11
13.1.3.	Climate and Acidification	13-12
Table 13-3 Summary on climate modification of acidification in Appendix 4
and Appendix 5 in addition to those in Appendix 13.	13-15
13.1.4.	Nitrogen, Climate, and Biodiversity	13-16
Table 13-4 Summary on climate modification of nitrogen (N) effects on
biodiversity in Appendix 6 in addition to those in Appendix 13.	13-19
13.2.	Estuaries	13-20
13.3.	Wetlands	13-21
APPENDIX 14. ECOSYSTEM SERVICES	14-1
14.1.	Introduction	14-1
14.2.	Ecosystem Services Frameworks	14-2
14.3.	United States Applications	14-4
14.3.1.	Acidification	14-5
Table 14-1 Ecosystem services research related to ecosystem
acidification.	14-7
14.3.2.	Eutrophication	14-8
Figure 14-1 Economic nitrogen cascade in the Chesapeake Bay
Watershed. 	14-11
Table 14-2 Potential damage costs of nitrogen (N) ($/kg N; 2008 or as
reported) to air, land, and water resources in the conterminous
U.S. in the early 2000s as synthesized by Sobota et al. (2015).
Low, median, and high costs derive from the specific damage
cost reference. Negative values indicate an economic benefit.	14-12
Figure 14-2 Map of ecosystem services altered by nitrogen critical load
exceedance. 	14-15
Table 14-3 Numbers of chains, Final Ecosystem Goods and Services
(FEGS), and beneficiaries (bens) associated with each initial
biological indicator Clark et al. (2017).	14-16
Table 14-4 Ecosystem services research related to Nitrogen (N)-driven
eutrophication.	14-17
14.3.3.	Nitrogen and Climate Modification 	14-17
June 2018	xxiv	DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Table 14-5 Summary of recent literature examining economic impacts of
ocean acidification on U.S. fisheries.	14-19
14.4.	European and Canadian Applications	14-19
Figure 14-3 Benefits and costs associated with the 25% decline in nitrogen
deposition in the U.K. since 1990.	14-21
14.5.	Global Perspective	14-21
14.6.	Summary	14-22
14.7.	Supplemental Materials: Ecosystem Services Profiles of Select Species	14-24
14.7.1.	Balsam Fir	14-24
14.7.2.	Eel Grass	14-25
14.7.3.	Green Turtle	14-27
14.7.4.	White Ash	14-28
14.7.5.	Lace Lichen	14-30
APPENDIX 15. OTHER ECOLOGICAL EFFECTS OF PARTICULATE MATTER	15-1
15.1.	Introduction	15-1
15.2.	Particulate Matter Deposition to Ecosystems	15-3
15.2.1.	Metals	15-3
15.2.2.	Organics	15-4
15.3.	Effects of Particulate Matter on Vegetation	15-5
15.3.1.	Vegetative Surfaces	15-5
15.3.2.	Foliar Uptake of Particulate Matter	15-7
15.3.3.	Particulate Matter Impacts on Gas Exchange Processes	15-8
15.3.4.	Plant Physiology	15-8
15.3.5.	Uptake of Particulate Matter by Plants from Soils	15-9
15.3.6.	Effects on Plant Growth and Reproduction	15-10
15.4.	Effects of Particulate Matter on the Soil Environment	15-11
15.4.1.	Bioavailability in Soils	15-11
15.4.2.	Soil Nutrient Cycling	15-12
15.4.3.	Soil Community Effects	15-14
15.4.4.	Soil Microbe Interactions with Plant Uptake of Particulate Matter	15-17
15.4.5.	Effects of Particulate Matter on Physical Properties of Soils	15-18
15.5.	Effects of Particulate Matter on Fauna	15-19
15.5.1.	Laboratory Bioassays	15-19
15.5.2.	Wildlife as Biomonitors of Particulate Matter	15-21
15.5.3.	Biomagnification	15-22
15.6.	Effects of Particulate Matter on Ecological Communities and Ecosystems	15-23
15.6.1.	Gradient Effects near Smelters	15-24
15.6.2.	Urban Environments	15-25
15.6.3.	Aquatic Ecosystems	15-25
15.6.4.	Experimental Microecosystems	15-26
15.7.	Summary of Ecological Effects of Particulate Matter	15-26
APPENDIX 16. CASE STUDIES 	16-1
Figure 16-1 CL exceedance in Class I areas.	16-1
16.1. Northeastern U.S. Case Study: Acadia National Park, Hubbard Brook Experimental
Forest, and Bear Brook Watershed	16-2
16.1.1. Background	16-2
Figure 16-2 Locations of northeastern U.S. case study areas and nearby
human population centers. 	16-3
Table 16-1 Selected characteristics of northeastern case study areas.	16-3
June 2018
XXV
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 16-3 Site map of Hubbard Brook Experimental Forest in the White
Mountains of New Hampshire.	
Table 16-2
Figure 16-4
Table 16-3
Table 16-4
Land use/land cover for northeastern case study areas. _
Land cover in the Northeast case study region.	
Literature cited by Northeast U.S. case study area.	
16.1.2. Deposition_
Figure 16-5
Key recent research literature focused on the case study
region.	
Total nitrogen deposition (A) and percentage of oxidized
nitrogen deposition (B) for the Northeast case study area
estimated by the National Atmospheric Deposition Program
Total Deposition Science committee.	
Figure 16-6 Total sulfur deposition (A) for the Northeast case study area
estimated by the National Atmospheric Deposition Program
Total Deposition Science committee. Time series of wet
deposition (B) from the National Atmospheric Deposition
Program/National Trends Network in the Hubbard Brook
Experimental Forest, NH. 	
16.1.3.
Critical Loads and Other Dose-Response Relationships 	
Table 16-5 Terrestrial empirical and modeling research on the response of
Table 16-6
nitrogen and sulfur deposition for the northeastern U.S.	
Empirical critical loads for nitrogen in Acadia National Park,
receptor, from Pardo et al. (2011c).	
by
Figure 16-7 Annual stream calcium and magnesium export (paired bars),
and cumulative excess export in West Bear Brook compared to
East Bear Brook (line), over the study period 1989-2000 at the
Bear Brook Watershed experiment.	
Table 16-7
Figure 16-8
Table 16-8
Table 16-9
Critical loads of nutrient nitrogen for the Northern Forests
ecoregion.	
Conceptual diagram showing the direct and indirect effects of
changes in temperature and precipitation on biogeochemical
processes in forests and on the services forests provide. Also
shown are feedbacks that further influence climatological
effects.	
Aquatic empirical research on the response of nitrogen and
sulfur deposition for the northeastern U.S. 	
Critical and target load and exceedance modeling studies in
the northeastern U.S.
Table 16-10 Empirical and modeled nitrogen critical loads applicable to the
northeastern U.S.
16.1.4.
Long-Term Ecological Monitoring	
Table 16-11 Summary table of observed terrestrial and aquatic acidification
long-term trends in Hubbard Brook Experimental Forest and
Bear Brook Watershed.
Table 16-12 Example surface water acidification chemistry studies in the
Northeast case study region.	
Table 16-13 Summary of observed terrestrial and aquatic nutrient
enrichment long-term responses in Acadia National Park,
Hubbard Brook Experimental Forest, Bear Brook Watershed,
and other northeastern regions.	
16.1.5. Recovery _
16.2. Adirondack Case Study: Adirondack Region of New York
16.2.1.
16.2.2.
Background_
Figure 16-9
Deposition_
Map of Adirondack park land cover.
16-5
16-8
16-8
16-9
16-12
16-13
16-14
16-15
16-16
16-17
16-20
16-22
16-23
16-25
16-26
16-30
16-32
16-33
16-34
16-37
16-40
16-44
16-45
16-45
16-47
16-48
June 2018
xxvi
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 16-10 Total nitrogen deposition (A) and percentage of oxidized
nitrogen deposition (B) for the Adirondack case study area
estimated by the National Atmospheric Deposition Program
Total Deposition Science committee.	16-49
Figure 16-11 Total sulfur deposition (A) for the Adirondack case study area
estimated by the National Atmospheric Deposition Program
Total Deposition Science committee. Time series of wet
deposition (B) from the National Atmospheric Deposition
Program/National Trends Network Whiteface Mountain, NY	16-50
16.2.3.	Critical Loads and Other Dose-Response Relationships 	16-51
Table 16-14 Terrestrial empirical and modeling research on the response of
nitrogen and sulfur deposition for the Adirondack Mountains.	16-51
Figure 16-12 Number offish species per lake verses acidity status,
expressed as acid neutralizing capacity, for Adirondack lakes.	16-55
Figure 16-13 Total macroinvertebrate species (community) richness as a
function of median pH in 36 streams sampled in the western
Adirondack Mountains of New York, 2003-2005; the four
standard (New York State) impact categories for species
richness are defined.	16-57
Figure 16-14 Species richness of biotic groups in 30 Adirondack study lakes
relative to midsummer epilimnetic pH during sample years.	16-58
Table 16-15 Aquatic empirical research on the response of nitrogen and
sulfur deposition for the Adirondacks.	16-59
Table 16-16 Critical and target load and exceedance modeling studies in
Adirondack Mountains. 	16-61
16.2.4.	Long-Term Monitoring	16-62
Figure 16-15 (A) Midsummer sulfate concentration in the epilimnion of
Brooktrout Lake (o) and in annual wet deposition (•) at local
National Atmospheric Deposition Program/National Trends
Network Station NY52 from 1984-2012. (B) Midsummer
epilimnetic concentrations of inorganic monomeric aluminum
(~) and hydrogen ion (¦) in Brooktrout Lake from 1984-2012.
(C) Midsummer phytoplankton (A) and total plankton
(phytoplankton, rotifers, crustaceans) (A) species richness in
Brooktrout Lake from 1984-2012.	16-67
16.3. Southeastern Appalachia Case Study	16-68
16.3.1.	Background	16-68
Table 16-17 Species in the Southeast case study region that are listed as
threatened or endangered or as a species of concern.	16-71
Figure 16-16. Great Smoky Mountains National Park and nearby Class I
wilderness areas, with emphasis on water sampling locations
within GSMNP and critical loads for watersheds described in
Appendix 16.2.3.2.	16-76
Figure 16-17 Land cover in the southern Appalachian Mountains case study
region.	16-77
16.3.2.	Deposition	16-77
Figure 16-18 Deposition over Great Smoky Mountain National Park.	16-78
Figure 16-19 Trends in wet deposition of nitrogen and sulfur in Great Smoky
Mountain National Park, 1990-2014. 	16-79
Figure 16-20 Maps showing total nitrogen deposition on left, total sulfur
deposition on right, for the 3-year average, 2011-2013.	16-80
Figure 16-21 Modeled sulfur and nitrogen deposition to the Great Smoky
Mountain National Park for the Year 2000.	16-81
Figure 16-22. Annual-averaged monthly footprint of reactive N deposition in
Great Smoky Mountain National Park (10.4 kg N/ha/yr), and
pie chart of fractional contribution from emission sectors, as
estimated by GEOS-Chem adjoint model.	16-82
16.3.3.	Critical Loads and Other Dose-Response Relationships 	16-82
June 2018
xxvii
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
16.3.4. Characterization and Long-Term Monitoring	16-85
Table 16-18 Example soil, terrestrial biota, and surface water acidification
characterization and long-term monitoring studies in the
southern Appalachian Mountains region.	16-86
16.4.	Tampa Bay Case Study	16-89
16.4.1.	Background	16-89
Figure 16-23 Tampa Bay mainstem segments and watershed. Also shown
are the National Atmospheric Deposition (NADP) National
Trends Network wet deposition monitoring site FL41 atVerna
wellfield in Sarasota County and the NADP Atmospheric
Integrated Research Monitoring Network site FL18 in
Hillsborough County.	16-91
Figure 16-24 Tampa Bay overview map highlighting watershed development
and land use.	16-94
16.4.2.	Deposition	16-95
Figure 16-25 Annual wet deposition of ammonium, nitrate, sulfate, and
acidity at the National Atmospheric Deposition National Trends
Network monitoring site closest to the Tampa Bay case study
area. 	16-95
Figure 16-26 Estimated annual loads of total nitrogen from various sources
to Tampa Bay summarized from 1976 to 2011.	16-96
Figure 16-27 (A) Wet and dry nitrogen deposition in Tampa Bay and the
surrounding area. (B) Percentage of oxidized nitrogen
deposition in Tampa Bay and the surrounding area.	16-98
16.4.3.	Long-Term Ecological Monitoring	16-98
Figure 16-28 Seagrass Loss in Tampa Bay.	16-100
Figure 16-29 Comparison of observed chlorophyll a and that predicted from
the total nitrogen load—chlorophyll a relationships for all four
mainstem Tampa Bay segments, for 1986-1998 and
1999-2007.	16-101
Figure 16-30 Total seagrass coverage in Tampa Bay circa 1950 through
2014.	16-103
Figure 16-31 Trend in hydrologically normalized total nitrogen load to Tampa
Bay relative to population increases in the Tampa Bay
metropolitan area.	16-104
16.4.4.	Nitrogen Man16-agement	16-105
Table 16-19 Numeric nutrient criteria for chlorophyll a for the four mainstem
segments of Tampa Bay adopted by the Florida Department of
Environmental Protection.	16-105
Table 16-20 Numeric nutrient criteria for total nitrogen for the four mainstem
segments of Tampa Bay.	16-106
16.5.	Rocky Mountain National Park Case Study	16-108
16.5.1.	Background	16-108
Figure 16-32 Rocky Mountain National Park ecosystems.	16-109
Figure 16-33 Rocky Mountain National Park land coverage using the land
cover classifications as mapped by the National Land Cover
Dataset. Percentage of cover is shown for the four dominant
cover types.	16-112
Figure 16-34 Rocky Mountain National Park hydrologic unit code 12
watersheds.	16-113
16.5.2.	Deposition	16-114
Figure 16-35 Spatial patterns of atmospheric nitrogen and sulfur deposition
in the Rocky Mountain National Park region based on TDEP
calculations averaged from 2011 -2013 (see Appendix 2) and
long-term trends in wet atmospheric deposition from the
Beaver Meadows National Atmospheric Deposition Program
Monitoring site within Rocky Mountain National Park.	16-115
June 2018
xxviii
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 16-36 Spatial patterns in the percentage of atmospheric nitrogen
deposited in oxidized forms and through wet deposition in the
Rocky Mountain National Park region based on TDEP
calculations averaged from 2011-2013 (see Appendix 2).	16-116
Figure 16-37 Rocky Mountain National Park nitrogen cycle. 	16-117
Figure 16-38 Total wet and dry deposition of nitrogen components measured
at Rocky Mountain National Park from November 2008 to
November 2009, including organic nitrogen and particulate
organic nitrogen.	16-118
Figure 16-39 Annual-averaged monthly footprint of reactive N deposition in
Rocky Mountain National Park(4.0 kg N/ha/yr), and pie chart of
fractional contribution from emission sectors, as estimated by
GEOS-Chem adjoint model. 	16-119
16.5.3.	Critical Loads and Other Dose-Response Relationships 	16-119
Table 16-21 Terrestrial empirical critical loads of nutrient nitrogen for the
Northwestern Forested Mountains ecoregion.	16-120
Table 16-22 Montane forest and alpine ecosystem critical loads for nitrogen
deposition research published since the critical load
assessment by Pardo et al. (2011 c).	16-123
Table 16-23 Hindcast absolute and percentage changes in species
abundance between 1900 and 2010 in response to historical
reconstructions of nitrogen deposition (Sullivan et al., 2005)
and historical climate change (IPCC, 2007b).	16-126
Table 16-24 Critical loads for nitrogen for eutrophication for surface water
(high-elevation lakes) in the Rocky Mountains.	16-128
Table 16-25 Lake water nitrate concentrations in nitrogen deposition studies
observing phytoplankton responses.	16-129
Table 16-26 Paleolimnological biological responses in Rocky Mountain
lakes exposed to anthropogenic nitrogen deposition.	16-130
Table 16-27 Critical loads of nitrogen or sulfur for surface water acidification
Rocky Mountain National Park and other high-elevation lakes
in the Rocky Mountains. 	16-132
Figure 16-40 The continuum of ecological sensitivity to nitrogen deposition.	16-134
Figure 16-41 Critical load thresholds for current and possible future
biogeochemical and biological effects of nitrogen deposition.	16-135
16.5.4.	Highlights of Additional Research Literature and Federal Reports since
January 2008 	16-135
Table 16-28 Summary of freshwater eutrophication studies in the Rocky
Mountains since 2008.	16-137
16.5.5.	Rocky Mountain National Park Initiative	16-138
Figure 16-42 Rocky Mountain National Park Initiative glidepath and current
wet nitrogen deposition at Loch Vale in Rocky Mountain
National Park.	16-140
Figure 16-43 Rocky Mountain National Park Initiative accomplishment
timeline.	16-141
16.5.6.	Interactions between Nitrogen Deposition, Precipitation, and Large-Scale
Ecological Disturbances	16-141
16.6. Southern California	16-143
16.6.1.	Background	16-143
Figure 16-44 Map of the distribution of vegetation types and land cover in
California. 	16-144
Table 16-29 Land coverages of Sequoia, Kings Canyons, and Joshua Tree
national parks.	16-147
Figure 16-45 Southern California case study region showing locations of
human population centers. 	16-148
16.6.2.	Deposition	16-149
June 2018
xxix
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 16-46 Patterns and temporal trends of nitrogen and sulfur deposition
in Joshua Tree National Park and surrounding region in
California. A and B show the 3-year average total deposition of
nitrogen and sulfur for 2011-2013	16-150
Figure 16-47 Patterns and temporal trends of nitrogen and sulfur deposition
in Joshua Tree National Park and surrounding region in
California. A shows the partitioning between oxidized and
reduced nitrogen; B and C show the 3-year average total
percentage of wet deposition of nitrogen and sulfur for
2011-2013.	16-151
Figure 16-48 Patterns and temporal trends of nitrogen and sulfur deposition
in Joshua Tree National Park and surrounding region in
California. The 25-year time series for wet deposition of nitrate,
ammonium, sulfate, and hydrogen ion obtained from the
National Atmospheric Deposition Program/National Trends
Network monitoring site CA 67.	16-152
Figure 16-49 Patterns and temporal trends of nitrogen and sulfur deposition
of Sequoia and Kings Canyons national parks and surrounding
region in California. A and B show the 3-year average total
deposition of nitrogen and sulfur for 2011-2013.	16-153
Figure 16-50 Patterns and temporal trends of nitrogen and sulfur deposition
of Sequoia and Kings Canyons national parks and surrounding
region in California. A shows the partitioning between oxidized
and reduced nitrogen, indicated as the fraction of total nitrogen
which is oxidized; B and C show the 3-year average total
percentage of wet deposition of nitrogen and sulfur for
2011-2013.	16-154
Figure 16-51 Patterns and temporal trends of nitrogen and sulfur deposition
of Sequoia and Kings Canyons national parks and surrounding
region in California. A and B show the 25-year time series for
wet deposition of nitrate, ammonium, sulfate, and hydrogen
obtained from the National Atmospheric Deposition
Program/National Trends Network monitoring sites CA99 and
CA75.	16-155
Figure 16-52 Annual-averaged monthly footprint of reactive N deposition in
Joshua Tree National Park (3.2 kg N/ha/yr) and Sequoia
National Park (5.7 kg N/ha/yr). Also shown for each park is a
pie chart of fractional contribution from emission sectors, as
estimated by GEOS-Chem adjoint model.	16-157
16.6.3.	Critical Loads and Other Dose-Response Relationships 	16-157
Figure 16-53 Composite critical load exceedance maps for all seven
vegetation types included in the study of Fenn et al. (2010)
showing the combined exceedance areas and the level of
exceedance (kg N/ha/yr).	16-160
Table 16-30 Summary of recent empirical dose-response and critical load
studies focused on the southern California case study area and
published since Pardo et al. (2011c).	16-161
Table 16-31 Terrestrial critical and target load and exceedance modeling
studies in southern California.	16-164
Table 16-32 Example surface water acidification studies in Sequoia and
Kings Canyons National Parks and other Sequoia and Kings
Canyons National Parks—relevant areas in the southern
California case study region.	16-168
16.6.4.	Highlights of Additional Research Literature and Federal Reports since
January 2008 	16-172
Table 16-33 Key recent research literature focused on the case study
region.	16-173
16.6.5.	Summary	16-174
June 2018
XXX
DRAFT: Do Not Cite or Quote

-------
CONTENTS (Continued)
Figure 16-54 Continuum of critical loads in southern California case study
area and relevant surrounding region.	16-175
REFERENCES	R-1
June 2018
xxxi
DRAFT: Do Not Cite or Quote

-------
INTEGRATED SCIENCE ASSESSMENT TEAM FOR
OXIDES OF NITROGEN, OXIDES OF SULFUR, AND
PARTICULATE MATTER—ECOLOGICAL CRITERIA
Executive Direction
Dr. John Vandenberg (Director, RTP Division)—National Center for Environmental
Assessment—RTP Division, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Dr. Steven J. Dutton (Deputy Director, RTP Division)—National Center for Environmental
Assessment, Office of Research and Development, U.S. Environmental Protection
Agency, Research Triangle Park, NC
Dr. Andrew Hotchkiss (Acting Branch Chief)—National Center for Environmental
Assessment, Office of Research and Development, U.S. Environmental Protection
Agency, Research Triangle Park, NC
Scientific Staff
Dr. Tara Greaver (Assessment Lead)—National Center for Environmental Assessment,
Office of Research and Development, U.S. Environmental Protection Agency, Research
Triangle Park, NC
Dr. Emmi Felker-Quinn—National Center for Environmental Assessment, Office of
Research and Development, U.S. Environmental Protection Agency, Research Triangle
Park, NC
Dr. Jeffrey D. Herrick—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Meredith Lassiter—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Stephen D LeDuc—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Washington, DC
Dr. Stephen McDow—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Technical Support Staff
Ms. Marieka Boyd—National Center for Environmental Assessment, Office of Research and
Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Mr. Kenneth J. Breito—Senior Environmental Employment Program, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Ms. Eleanor Jamison—Senior Environmental Employment Program, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
June 2018
xxxii
DRAFT: Do Not Cite or Quote

-------
Mr. Ryan Jones—National Center for Environmental Assessment, Office of Research and
Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Ms. Connie Meacham—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Ms. Olivia Philpott—Senior Environmental Employment Program, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Mr. Samuel S. Thacker—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Mr. Richard N. Wilson—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
June 2018
xxxiii
DRAFT: Do Not Cite or Quote

-------
AUTHORS, CONTRIBUTORS, AND REVIEWERS
Authors
Dr. Tara Greaver (Assessment Lead)—National Center for Environmental Assessment,
Office of Research and Development, U.S. Environmental Protection Agency, Research
Triangle Park, NC
Ms. Marion Deerhake*—RTI International, Research Triangle Park, NC
Dr. Emmi Felker-Quinn—National Center for Environmental Assessment, Office of
Research and Development, U.S. Environmental Protection Agency, Research Triangle
Park, NC
Dr. Jeffrey D. Herrick—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Meredith Lassiter—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Stephen D LeDuc—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Washington, DC
Dr. Stephen McDow—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Jennifer Phelan*—RTI International, Research Triangle Park, NC
Dr. Robert Pinder—Health and Environmental Impacts Division, Office of Air and
Radiation, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Joseph P. Pinto—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Timothy J. Sullivan*—E&S Environmental Chemistry, Inc., Corvallis, OR
Dr. Alan Talhelm—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
* Under contract with RTI International
Contributors
Mr. Adam Benson—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Dr. Mary C. Barber*—RTI International, Washington, DC
Ms. Tamara Blett—National Park Service, Lakewood, CO
Dr. Jana Compton—National Health and Environmental Effects Research Laboratory, Office
of Research and Development, U.S. Environmental Protection Agency, Corvallis, OR
Dr. Christopher M. Clark—National Center for Environmental Assessment, Office of
Research and Development, U.S. Environmental Protection Agency, Washington, DC
June 2018
xxxiv
DRAFT: Do Not Cite or Quote

-------
Dr. Linda Geiser—Air Resource Management National Program Leader, USDA Forest
Service, Washington, DC
Dr. Alan Knapp—Department of Biology, Colorado State University, Fort Collins CO
Dr. Jason Lynch—Office of Air and Radiation, Office of Atmospheric Programs, U.S.
Environmental Protection Agency, Washington, DC
Ms. April Maxwell—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Ms. Margaret O'Neil*—RTI International, Research Triangle Park, NC
Ms. Jennifer Richkus*—RTI International, Research Triangle Park, NC
Ms. Elizabeth Sullivan*—RTI International, Research Triangle Park, NC
Dr. George L. Van Houtven*—RTI International, Research Triangle Park, NC
* Under contract with RTI International
Reviewers
Mr. Barry P. Baldigo, Water Science Center, U.S. Geological Survey, Troy, NY
Ms. Tamara Blett—National Park Service, Lakewood, CO
Dr. Michael Bell—National Park Service, Lakewood, CO
Dr. Micah Bennett—National Center for Environmental Assessment, Office of Research and
Development, U.S. Environmental Protection Agency, Washington, DC
Ms. Patricia Brewer—National Park Service, Lakewood, CO
Dr. Suzanne Bricker—National Oceanic and Atmospheric Administration, Washington, DC
Mr. Jim Cheatham—National Park Service, Lakewood, CO
Dr. Christopher M. Clark—National Center for Environmental Assessment, Office of
Research and Development, U.S. Environmental Protection Agency, Washington, DC
Dr. Christopher B. Craft—Indiana University, Bloomington, IN
Dr. Eric Davidson—University of Maryland, Appalachian Laboratory, Frostburg, MD
Ms. Christine Davis—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Jean-Jacques Dubois—North Carolina State University, Raleigh, NC
Dr. Eric Edgerton—Atmospheric Research and Analysis, Cary, NC
Ms. Candace Edwards—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Ms. Charlene Finley—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
June 2018
XXXV
DRAFT: Do Not Cite or Quote

-------
Ms. Elizabeth Gatling—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Dr. Cynthia Gilmour—Smithsonian Environmental Research Center, Edgewater, MD
Mr. William Grffin—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Cinncinnatti, OH
Dr. David Grantz—College of Natural and Agricultural Sciences, Air Pollution Research
Center, University of California Riverside, Parlier, CA
Dr. Michael Griffith—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Washington, DC
Dr. Scot Haggerthey—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Washington, DC
Dr. Meredith K. Hastings—Earth, Environmental, and Planetary Sciences, Brown
University, Providence, RI
Dr. Laszlo Horvath—Plant Ecology Research Group, Szent Istvan University, Hungary
Dr. Robert Hetes—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
Mr. James Hou—Office of the Regional Administrator, Region 8, U.S. Environmental
Protection Agency, Denver, CO
Dr. Bryan Hubbel—Office of Air Quality Planning and Standards, Office of Air and
Radiation, U.S. Environmental Protection Agency, Research Triangle Park, NC
Mr. Saturo Ito—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Mr. Ryan Jones—National Center for Environmental Assessment, Office of Research and
Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. James Kaldy—National Health and Environmental Effects Research Laboratory, Office
of Research and Development, U.S. Environmental Protection Agency, Corvallis, OR
Mr. Matt Kulp—National Park Service, Gatlinburg, TN
Ms. Emily Lau—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Dr. Sylvia Lee—National Center for Environmental Assessment, Office of Research and
Development, U.S. Environmental Protection Agency, Washington, DC
Dr. Jason Lynch—Office of Air and Radiation, Office of Atmospheric Programs, U.S.
Environmental Protection Agency, Washington, DC
Mr. Ihab Mikati—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
June 2018
xxxvi
DRAFT: Do Not Cite or Quote

-------
Ms. Danielle Moore—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Ms. Kristi Morris—National Park Service, Lakewood, CO
Dr. Kristopher Novak—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Caroline Nowlan—Harvard-Smithsonian Center for Astrophysics, Harvard University,
Cambridge, MA
Mr. Kyle Painter—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Ms. Rebecca Perrin—Office of the Regional Administrator, Region 8, U.S. Environmental
Protection Agency, Denver, CO
Mr. Jim Renfro—National Park Service, Gatlinburg, TN
Dr. Jennifer Richmond-Bryant—National Center for Environmental Assessment, Office of
Research and Development, U.S. Environmental Protection Agency, Research Triangle
Park, NC
Dr. Caroline Ridley—National Center for Environmental Assessment, Office of Research
and Development, U.S. Environmental Protection Agency, Washington, DC
Ms. Kristin Riha—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
Ms. Alexandra Ross—Oak Ridge Institute for Science and Education, National Center for
Environmental Assessment, Office of Research and Development, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Dr. Donald Scavia—University of Michigan, Ann Arbor, MI
Ms. Vicki Sandiford—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Bret Schichtel—National Park Service, Fort Collins, CO
Dr. Erika Sasser—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. J. Travis Smith—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
Ms. Ginger Tennant—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Barbara J. Turpin—Gillings School of Global Environmental Health, University of
North Carolina, Chapel Hill, NC
Dr. Rich Scheffe—Office of Air Quality Planning and Standards, U.S. Environmental
Protection Agency, Research Triangle Park, NC
Mr. Randy Waite—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
June 2018
xxxvii
DRAFT: Do Not Cite or Quote

-------
Dr. John T. Walker—National Risk Management Research Laboratory, Office of Research
and Development, U.S. Environmental Protection Agency, Research Triangle Park, NC
Dr. Christopher Weaver—National Center for Environmental Assessment, Office of
Research and Development, U.S. Environmental Protection Agency, Washington, DC
Ms. Karen Wesson—Office of Air and Radiation, Office of Air Quality Planning and
Standards, U.S. Environmental Protection Agency, Research Triangle Park, NC
June 2018
xxxviii
DRAFT: Do Not Cite or Quote

-------
CLEAN AIR SCIENTIFIC ADVISORY COMMITTEE
OXIDES OF NITROGEN, OXIDES OF SULFUR, AND
PARTICULATE MATTER—ECOLOGICAL CRITERIA
NAAQS REVIEW PANEL
Chair of Review Panel
Dr. Ivan J. Fernandez**—Distinguished Maine Professor, School of Forest Resources and
Climate Change Institute, University of Maine, Orono, ME
Review Panel Members
Dr. Edith Allen—Professor of Plant Ecology, Department of Botany and Plant Sciences,
University of California Riverside, Riverside, CA
Dr. Praveen Amar—Independent Consultant, Environment, Energy, and Climate Strategies,
Lexington, MA
Dr. James Boyd—Senior Fellow and Director, Center for the Management of Ecological
Wealth, Resources for the Future, Washington, DC
Dr. Elizabeth W. Boyer—Associate Professor of Water Resources, Department of
Ecosystem Science and Management, Pennsylvania State University, University Park, PA
Dr. Douglas Burns—Research Hydrologist, New York Water Science Center, U.S.
Geological Survey, Troy, NY
Ms. Lauraine Chestnut—Managing Economist, Stratus Consulting Inc., Boulder, CO
Dr. Charles T. Driscoll, Jr.—Distinguished Professor and University Professor of
Environmental Systems Engineering, Department of Civil and Environmental
Engineering, College of Engineering and Computer Science, Syracuse University,
Syracuse, NY
Dr. Mark Fenn—Research Plant Pathologist, Pacific Southwest Research Station, USDA
Forest Service, Riverside, CA
Dr. James Galloway—Sidman P. Poole Professor of Environmental Sciences, Department of
Environmental Sciences, University of Virginia, Charlottesville, VA
Dr. Frank Gilliam, Professor—Department of Biological Sciences, Marshall University,
Huntington, WV
Dr. Robert A. Goldstein—Senior Technical Executive for Water and Ecosystems, Electric
Power Research Institute, Palo Alto, CA
Dr. Daven Henze—Assistant Professor and Charles C. Gates Faculty Fellow, Department of
Mechanical Engineering, University of Colorado, Boulder, CO
Dr. Robert W. Howarth—David R. Atkinson Professor of Ecology & Environmental
Biology, Department of Ecology and Evolutionary Biology, Cornell University, Ithaca,
NY
Dr. Donna Kenski—Data Analysis Director, Lake Michigan Air Directors Consortium,
Rosemont, IL
June 2018
xxxix
DRAFT: Do Not Cite or Quote

-------
Dr. William McDowell—Professor of Environmental Science, Department of Natural
Resources and the Environment, University of New Hampshire, Durham, NH
Dr. Erik Nelson—Assistant Professor, Department of Economics, Bowdoin College,
Brunswick, ME
Dr. Hans Paerl—Kenan Professor of Marine and Environmental Sciences, Institute of
Marine Sciences, University of North Carolina—Chapel Hill, Morehead City, NC
Mr. Richard L. Poirot—Air Quality Planning Chief, Air Quality and Climate Division,
Vermont Department of Environmental Conservation, Montpelier, VT
Dr. Armistead (Ted) Russell—Howard T. Tellepsen Chair and Regents Professor of Civil
and Environmental Engineering, Department of Civil and Environmental Engineering,
Georgia Institute of Technology, Atlanta, GA
Dr. Stephen E. Schwartz—Senior Scientist, Environmental and Climate Sciences
Department, Brookhaven National Laboratory, Upton, NY
Dr. Kathleen Weathers—Senior Scientist, Cary Institute of Ecosystem Studies, Millbrook,
NY
**Member of the statutory CASAC appointed by the U.S. EPA Administrator
Science Advisory Board Staff
Dr. Thomas Armitage—Designated Federal Officer, U.S. Environmental Protection Agency,
Science Advisory Board (1400R), 1200 Pennsylvania Avenue, NW, Washington, D.C.
20460-0001, Phone: 202-564-2155, Fax: 202-565-2098, (Armitage.thomas@epa.gov)
June 2018
xl
DRAFT: Do Not Cite or Quote

-------
ACRONYMS AND ABBREVIATIONS
Acronym/Abbreviation Meaning
ACCENT	Atmospheric Composition
Change: the European Network
of excellence
AIRMoN	Atmospheric Integrated
Research Monitoring Network
AIRS	Atmospheric Infrared Sounder
(instrument)
A1	aluminum
Al3+	aluminum ion
Ali	inorganic aluminum
Al2+	aluminum ion
Alo	organic aluminum
Al(OH)3	aluminum hydroxide
ALSC	Adirondack Lake Survey
Corporation
ALTM	Adirondack Long Term
Monitoring
AMD	acid mine drainage
ANC	acid neutralizing capacity
ANPP	aboveground net primary
production
AOD	aerosol optical depth
AOSR	Athabasca oil sands region
AQCD	Air Quality Criteria Document
AQEG	Air Quality Expert Group
AQI	Air Quality Index
AQS	Air Quality System (database)
Ar	argon
ARS	Agricultural Research Service
As	arsenic
ASI	Acid Stress Index
asl	above sea level
ATMOS	Atmospheric Trace Molecule
Spectroscopy
ATTILA	type of Lagrangian model
AUSPEX	Atmospheric Utility Signatures,
Predictions, and Experiments
AVIRIS	Airborne Visible and Infrared
Imaging Spectrometer
Acronym/Abbreviation Meaning
Ba	barium
BBW	Bear Brook Watershed
BBWM	Bear Brook Watershed, Maine
Be or BC	base cation
BCE	exchangeable base cations
BCS	base-cation surplus
BCw	base cation weathering
BGC	Biogeochemical (model)
B-IBI	benthic index of biological
integrity
BMPs	best management practices
BNF	biological nitrogen fertilization
Br	bromine
Br~	bromide ion
Bn	molecular bromine
BrCl	bromine chloride
BrO	bromine mooxide
BUV	Backscatter Ultraviolet
Spectrometer
BUVD	Beneficial Use Values Database
C	carbon; concentration
12C	carbon-12, stable isotope of
carbon
13C	carbon-13, stable isotope of
carbon
Ca	ambient air concentration
Ca	calcium
Ca2+	calcium ion
CAA	Clean Air Act
CAAA	Amendments to the Clean Air
Act
CAAAC	Clean Air Act Advisory
Committee
CaCk	calcium chloride
CaCC>3	calcium carbonate
CALIPSO	Cloud-Aerosol Lidar and
Infrared Pathfinder Satellite
Observation (satellite)
June 2018
xli
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation
Ca(N03)2
Ca(OH)2
CAPMoN
CaS04-2H20
CASTNet
CB4
Cd
CEC
CENTURY
CFCs
CG
Chi a
CH4
C2H4
C2H6
C5H8
CH3CHO
CH3C(0)
CH3C(0)00
CH2I2
CH-20
CH300H
CH3-S-CH3
CH3-S-H
(CH3)2SO
CH3S03H
CH3-S-S-CH3
Ci
CL
CI
cr
C12
CLaMS
Meaning
calcium nitrate
calcium hydroxide
Canadian Air and Precipitation
Monitoring Network
gypsum
Clean Air Status and Trends
Network
Carbon Bond 4 (model)
cadmium
cation exchange capacity
model that simulates carbon,
nitrogen, phosphorus, sulfur, and
water dynamics in the soil-plant
system at monthly intervals over
time scales of centuries and
millennia
chlorinated fluorocarbons
cloud-to-ground (lightning flash)
chlorophyll a
methane
ethene
ethane
isoprene
acetaldehyde
acetyl radical
acetyl peroxy radical
diiodomethane
formaldehyde
methyl hydroperoxide
dimethylsulfide, DMS
methyl mercaptan
dimethyl sulfoxide, DMSO
methanesulfonic acid
dimethyl disulfide, DMDS
interstitial air concentration
critical load
chlorine
chloride ion
molecular chlorine
type of Lagrangian model
Acronym/Abbreviation
CloudSat
CINO2
CMAQ
CMSA
CO
CO2
CO3-
CONUS
CPUE
CRREL
CS
CS2
CSS
CTM
Cu
CV
CVM
A, 5
DayCent
DayCent-Chem
DC
DDRP
DDT
DECOMP
DEP
DIC
DIN
DMDS
DMS
DMSO
DNDC
Meaning
NASA Earth observation
satellite
nitryl chloride
Community Multiscale Air
Quality (modeling system)
consolidated metropolitan
statistical area
carbon monoxide
carbon dioxide
carbonate
continental U.S.
catch per unit effort
U.S. Army Cold Regions
Research and Engineering
Laboratory
Consumer surplus
carbon disulfide
coastal sage scrub (ecosystem)
chemical transport model
copper
contingent valuation
contingent valuation method
delta, difference; change
model for daily biogeochemistry
for forest, grassland, cropland,
and savanna systems
combination of DayCent-Chem
and PHREEQC models
dichotomous choice
Direct Delayed Response Project
Damage Delay Time
decomposition model based on
soil-plant system dynamics
Department of Environmental
Protection
dissolved inorganic carbon
dissolved inorganic nitrogen
dimethyl disulfide,
CH3-S-S-CH3
dimethyl sulfide, CH3-S-CH3
dimethylsulfoxide
Denitrification-Decomposition
(model)
June 2018
xlii
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation Meaning	Acronym/Abbreviation Meaning
DO
dissolved oxygen
FISH
Fish in Sensitive Habitats
DOC
dissolved organic carbon

(project)
DON
dissolved organic nitrogen
FLEXPART
type of Lagrangian model
EBB
East Bear Brook
ForSAFE
three-component model using
nitrogen, carbon cycling, and
EC
elemental carbon

soil chemistry
EEAs
Essential Ecological Attributes
FRM
Federal Reference Method
ELA
Experimental Lakes Area
FTIR
Fourier Transform Infrared


Spectroscopy
ELS
Eastern Lakes Survey


FW2
black carbon soot
EMAP
Environmental Monitoring and



Assessment Program
Fx
flux
EMEFS
Eulerian Model Evaluation Field
Study
YN2O5
uptake coefficient for N2O5 on
particles
EMEP
Co-operative Programme for
Monitoring and Evaluation of
GAW
Global Atmospheric Watch
(program)

the Long-range Transmission of
GCE
Goddard Cumulus Ensemble

Air Pollutants in Europe

(model)
EMF
ectomycorrhizal fungi
GDP
gross domestic product
EOS
Earth Observation System
GEOS
Goddard Earth Observing
U.S. EPA
U.S. Environmental Protection

System

Agency
GEOS-Chem
Goddard Earth Observing
eq
equivalents

System (with global chemistry

transport model)
Goddard Earth Observing
ER
ecosystem respiration
GEOS-IDAS
EPT
Ephemeroptera-Plecoptera-
Tricoptera (index)

System Data Assimilation
System
ERP
Episodic Response Project
GFED
Global Fire Emissions Database
ESA
European Space Agency
GHG
greenhouse gas
EVRI
Environmental Valuation
Reference Inventory
GOES
Geostationary Operational
Environmental Satellites
F
flux
GOME
Global Ozone Monitoring
F"
fluoride ion

Experiment
FAB
First-order Acidity Balance
GPP
gross primary productivity

model
gs
stomatal conductance
FACE
free-air CO2 enrichment (studies)
GtC
gigaton carbon
Fe
iron
Gton
gigaton
FeP04
iron phosphate
GWP
global warming potential
FeS
iron sulfide
H
hydrogen; hydrogen atom
F-factor
fraction of the change in mineral
2H
hydrogen-2, deuterium, stable

acid anions that is neutralized by

isotope of hydrogen
proton, hydrogen ion; relative

base cation release
H+
FHM
Forest Health Monitoring

acidity
FIA
Forest Inventory and Analysis
(program)
ha
HAPs
hectare
hazardous air pollutants
June 2018
xliii
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation Meaning
HBEF	Hubbard Brook Experimental
Forest
HBES	Hubbard Brook Ecosystem
Study
HBN	Hydro logic Benchmark Network
HC	hydrocarbon
HCHO	formaldehyde
HC1	hydrochloric acid
HCC>3~	bicarbonate
Hg	mercury
HNO2, HONO	nitrous acid
HNO3, HOONO	nitric acid
E1NC>4	pernitric acid
HO2	hydroperoxyl radical
H2O2	hydrogen peroxide
HO2NO2	peroxynitric acid
HOBr	hypobromous acid
HOC1	hypochlorous acid
HOX	hypohalous acid
HP	hedonic pricing
HSC>3~	bisulfate ion
HSC>4~	sulfuric acid ion
H2S	hydrogen sulfide
H2SO3	sulfurous acid
H2SO4	sulfuric acid
HUC-8s	8 digit Hydrologic Unit Codes
hv	photon with energy at
wavelength v
I	iodine
I2	molecular iodine
IA	Integrated Assessment
IADN	Integrated Atmospheric
Monitoring Deposition Network
IC	intracloud (lightning flash)
ILWAS	Integrated Lake-Watershed
Acidification Study
IPC	International Cooperative
Programme
IEc	Industrial Economicsym
IIASA	International Institute for
Applied Systems Analysis
Acronym/Abbreviation Meaning
IMPROVE
Interagency Monitoring of

Protected Visual Environments
INOs
iodine nitrate
INTEX-NA
Intercontinental Chemical

Transport Experiment—North

America
IO
iodine oxide
IPCC
Intergovernmental Panel on

Climate Change
IPCC-AR4
Intergovernmental Panel on

Climate Change 4th Assessment

Report
IPCC-TAR
Intergovernmental Panel on

Climate Change 3rd Assessment

Report
IQR
interquartile range
IR
infrared
ISA
Integrated Science Assessment
J
flux from a leaf, deposition flux

(g/cm/second)
JK
Joyce Kilmer
JPL
Jet Propulsion Laboratory
JRGCE
Jasper Ridge Global Climate

Change Experiment
K
potassium
K+
potassium ion
Ka
dissociation constant
Kb
dissociation constant
KH
Henry's Law constant in M/atm

(M-atnT1)
kmol
kilomole
KNOs
potassium nitrate
Kw
ion product of water
LAF
Lake Acidification and Fisheries
LAR
leaf-area ratio
LB
laboratory bioassay
LC0.01
lethal concentration at which

0.01% of exposed animals die
LD33
lethal dose at which 33% of

exposed animals die
LDH
lactic acid dehydrogenase
LG
Linville Gorge
June 2018
xliv
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation Meaning
LIDAR
Light Detection and Ranging

(remote sensing system)
LIF
laser-induced fluorescence
LIMS
Limb Infrared Monitor of the

Stratosphere
LMCO
Lacasse-like multicopper

oxidase
LOD
limit of detection
LP
long-path
LRTAP
Long Range Transport of Air

Pollution
LTER
Long-Term Ecological Research

(program)
LTM
Long-Term Monitoring (project)
M
air molecule
MA
Millennium Ecosystem

Assessment
MAGIC
Model of Acidification of

Groundwater in Catchments

(model)
MAHA
Mid-Atlantic Highlands

Assessment of streams
MAQSIP
Multiscale Air Quality

Simulation Platform (model)
MAT
moist acidic tundra
MAX-DO AS
multiple axis differential optical

absorption spectroscopy
MBC
microbial biomass carbon
MBL
marine boundary layer
MDN
Mercury Deposition Network
MeHg
Methylmercury
MEM
model ensemble mean
Heq
Microequivalent
Mg
Magnesium
Mg2+
magnesium ion
MIMS
membrane inlet mass

spectrometry
MM5
National Center for Atmospheric

Research/Penn State Mesoscale

Model, version 5
Mn
Manganese
MOBILE6
Highway Vehicle Emission

Factor Model
Acronym/Abbreviation Meaning
MODIS
Moderate Resolution Imaging

Spectroradiometer
MOPITT
Measurement of Pollution in the

Troposphere (satellite

instrument)
MOZAIC
Measurement of Ozone and

Water Vapor by Airbus

In-Service Aircraft
MOZART
Model for Ozone and Related

Chemical Tracers
MPAN
peroxymethacrylic nitrate
MPCA
Minnesota Pollution Control

Agency
MSA
metropolitan statistical area
Mt
million, or mega tons
N
nitrogen
N, n
number of observations
14N
nitrogen-14, stable isotope of

nitrogen
15N
nitrogen-15, stable isotope of

nitrogen
N2
molecular nitrogen; nonreactive

nitrogen
N14C
plant soil N and C cycling model
NA
not available; insufficient data
Na
sodium
Na+
sodium ion
NAAQS
National Ambient Air Quality

Standards
NaCl
sodium chloride
NADP
National Atmospheric

Deposition Program
Na2Mo04
sodium molybdate
NAMS
National Air Monitoring Stations
NANI
Net anthropogenic nitrogen

inputs
NAPAP
National Acid Precipitation

Assessment Program
NASQAN
National Stream Quality

Accounting Network
NARSTO
program formerly known as

North American Regional

Strategy for Atmospheric Ozone
NAS
National Academy of Sciences
June 2018
xlv
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation Meaning
NASA	National Aeronautics and Space
Administration
Na2SC>4	sodium sulfate
NASQAN	National Stream Quality
Accounting Network
NATTS	National Air Toxics Trends
(network)
NAWQA	National Water Quality
Assessment (program)
NCore	National Core Monitoring
Network
N-dep	nitrogen deposition
NEE	net ecosystem exchange
NEG/ECP	New England Governors and
Eastern Canadian Premiers
NEI	National Emissions Inventory
NEON	National Ecological Observatory
Network
NEP	net ecosystem productivity
N-fert	nitrogen-fertilization
N-fix	nitrogen-fixing vegetation
NFI	net factor income
NH3	ammonia
NH2	amino (chemical group)
NH4+	ammonium ion
NH4CI	ammonium chloride
NH4NO3	ammonium nitrate
(NH4)2S04	ammonium sulfate
NHx	category label for NH3 plus
NH4+
NHy	total reduced nitrogen
Ni	nickel
NILU	Norwegian Institute for Air
Research
NITREX	Nitrogen saturation Experiments
nitro-PAH	nitro-polycyclic aromatic
hydrocarbon
NLCD	National Land Cover Data
Nmin	nitrogen mineralization
NMOC	nonmethane organic compound
NO	nitric oxide
NO2	nitrogen dioxide
Acronym/Abbreviation Meaning
no2-
nitrite
no3-
nitrate
N20
nitrous oxide
N205
dinitrogen pentoxide
NOAA
U.S. National Oceanic and

Atmospheric Administration
NOAA-ARL
U.S. National Oceanic and

Atmospheric Administration Air

Resources Laboratory
NOAEL
no-observed-adverse-effect level
NOEC
no-observed-effect concentration
NOx
sum of NO and NO2
NOy
sum of NOx and NOz; odd

nitrogen species; total oxidized

nitrogen
NOz
sum of all inorganic and organic

reaction products of NOx

(HONO, HNO3, HNO4, organic

nitrates, particulate nitrate,

nitro-PAHs, etc.)
NPOESS
National Polar-orbiting

Operational Environmental

Satellite System
NPP
net primary production
NPS
National Park Service
Nr
reactive nitrogen
NRC
National Research Council
NS or n.s.
nonsignificant
NSF
National Science Foundation
NSS
National Stream Survey
Nss
nonsea salt
NSTC
National Science and

Technology Council
NSWS
National Surface Water Survey
NTN
National Trends Network
NuCM
nutrient cycling model
O2
molecular oxygen
O3
ozone
16o
oxygen-16, stable isotope of

oxygen
18o
oxygen-18, stable isotope of

oxygen
19Q
oxygen-19, radioactive isotope

of oxygen
June 2018
xlvi
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation Meaning
OC	organic carbon
O-CN	terrestrial biosphere model
OCO	Orbiting Carbon Observatory
OCS	carbonyl sulfide
O('D)	electronically excited oxygen
atom
OH	hydroxyl radical
OM	organic matter
OMI	Ozone Monitoring Instrument
0(3P)	ground-state oxygen atom
P	phosphorus
P, p	probability value
Pi	1 st percentile
P5	5th percentile
P95	95th percentile
P99	99th percentile
PAHs	polycyclic aromatic
hydrocarbons
PAMS	Photochemical Assessment
Monitoring Stations
PAN	peroxyacetyl nitrate
PANs	peroxyacyl nitrates
PARASOL	Polarization and Anisotropy of
Reflectances for Atmospheric
Sciences coupled with
Observations from a Lidar
(satellite instrument)
Pb	lead
PBL	planetary boundary layer
PC	payment card
PCBs	polychlorinated biphenyl
compounds
pH	relative acidity
P(HN03)	production of nitric acid
PHREEQC	model for soil and water
geochemical equilibrium
PIRLA	Paleocological Investigation of
Recent Lake Acidification
(projects)
pKa	dissociation constant
PM	particulate matter
Acronym/Abbreviation Meaning
PM2.5	particulate matter with
aerodynamic diameter of <2.5
(jm
PM10
particulate matter with

aerodynamic diameter <10 (jm
PMlO-2.5
particulate matter with

aerodynamic diameter between

10 and 2.5 (im
PM-CAMx
Comprehensive Air Quality

Model with extensions and with

particulate matter chemistry
PnET
Photosynthesis and

EvapoTranspiration (model)
PnET-BGC
Photosynthesis and

EvapoTranspiration-

Biogeochemical (model)
PnET-CN
Photosynthesis and

EvapoTranspiration model of C,

water, and N balances
PnET-N-DNDC
Photosynthesis and

EvapoTranspiration-

Denitrification-Decomposition

(model)
pN03"
particulate nitrate
P(03)
production of O3
PO4-, PO43-
phosphate
POPs
persistent organic pollutants
Ppb
parts per billion
PPN
peroxypropionyl nitrate
Ppt
parts per trillion
PRB
policy relevant background
PRE-STORM
Preliminary Regional

Experiment for STORM
PROFILE
model using soil mineralogy as

input
PS
producer surplus
pS042~
particulate sulfate
P(S042~)
production of sulfate
Q
flow rate; discharge
Q10
temperature coefficient
QAPP
Quality Assurance Project Plan
R
generic organic group attached

to a molecule
R2
coefficient of determination
r2
correlation coefficient
June 2018
xlvii
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation
Ra
Raboveground
Rautotrophic
Rb
Rc
RADM
RAMS
RAPS
RCOO-s
RC(0)00
RDT
REMAP
RH
RLTM
Rmicrobial
RMCC
RMSE
RO2
RONO2
RO2NO2
ROS
RP
RRx
RR
Rsoil
RuBisCO
s
S
32g
34S
35S
SAA
SAFE
Meaning
aerodynamic resistance
aboveground respiration
soil autotrophic respiration
boundary layer resistance
internal resistance
Regional Acid Deposition Model
Regional Atmospheric Modeling
System
Regional Air Pollution Study
strongly acidic organic anions
organic peroxy radical
Recovery Delay Time
Regional Environmental
Monitoring and Assessment
Program
relative humidity
Regional Long-Term Monitoring
microbial respiration
Research and Monitoring
Coordinating Committee
root mean squared error
organic peroxyl; organic peroxy
organic nitrate
peroxynitrate
rain on snow
revealed preferences
lognormal-transformed response
ratio
response ratio
total soil respiration
ribulose-1,5-bisphosphate
carboxylase/oxygenase
second
sulfur
sulfur-32, stable isotope of sulfur
sulfur-34, stable isotope of sulfur
sulfur-35, radioactive isotope of
sulfur
sum of mineral acid anion
concentrations
Soil Acidification in Forest
Ecosystems (model)
Acronym/Abbreviation Meaning
SAMAB
SAMI
SAO
SAPRAC
SBC
SBUV
SC
SCAQS
SCIAMACHY
Se
SEARCH
Si
SIP
SJAQS
SLA
SLAMS
SMART
SMB
SMBE
SO
502
503
SO32-
S042~
S2O
SOM
SONEX
Southern Appalachian Man and
the Biosphere (program)
Southern Appalachian
Mountains Initiative
Smithsonian Astrophysical
Observatory
Statewide Air Pollution Research
Center
sum of base cation
concentrations
Solar Backscatter Ultraviolet
Spectrometer
safe concentration
Southern California Air Quality
Study
Scanning Imaging Absorption
Spectrometer for Atmospheric
Chartography
selenium; standard error
Southeastern Aerosol Research
and Characterization Study
(monitoring program)
silicon
State Implementation Plan
San Joaquin Valley Air Quality
Study
specific leaf area
State and Local Air Monitoring
Stations
Simulation Model for
Acidification's Regional Trends
(model)
Simple Mass Balance (model)
steady-state mass-balance
equations
sulfur monoxide
sulfur dioxide
sulfur trioxide
sulfite
sulfate ion
disulfur monoxide
soil organic matter
Subsonics Assessment Ozone
and Nitrogen Oxides Experiment
June 2018
xlviii
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation Meaning
SOS
Southern Oxidant Study
SOS/T
State of Science/Technology

(report)
SOx
sulfur oxides
SP
stated preferences
SPARROW
SPAtially Referenced

Regressions on Watershed

Attributes (model)
SR
Shining Rock
Sr
strontium
86 Sr
strontium-86, stable isotope of

strontium
87Sr
strontium-87, stable isotope of

strontium
SRB
sulfate-reducing bacteria
SRP
soluble reactive phosphorus
SSWC
Steady State Water Chemistry

(model)
STA
Soil Texture Approximations

(model)
STE
strata spheric-tropo spheric

exchange
STN
Speciation Trends Network
SUM06
seasonal sum of all hourly

average concentrations

>0.06 ppm
SVOC
semivolatile organic compound
SWAS
Shenandoah Watershed Study
T, T
tau, atmospheric lifetime
T
time; duration of exposure
TAF
Tracking and Analysis

Framework (model)
Tair
air temperature
TAMM
Timber Assessment Market

Model
TAR
Third Assessment Report
TC
total carbon; travel cost
TCM
travel cost method
TDLAS
Tunable Diode Laser Absorption

Spectrometer
Tg
teragram
TIME
Temporally Integrated

Monitoring of Ecosystems

(program)
Acronym/Abbreviation Meaning
TN
total nitrogen
TOC
total organic carbon
TOMS
Total Ozone Mapping

Spectrometer
TOR
tropospheric ozone residual
TP
total phosphorus
TRACE-P
Transport and Chemical

Evolution over the Pacific
TSI
timber-stand improvement
TSS
total suspended solids
Twater
water temperature
UAN
urea and ammonium nitrate

fertilizer
UMD-CTM
University of Maryland

Chemical Transport Model
UNECE
United Nations Economic

Commission for Europe
USDA
U.S. Department of Agriculture
USFS
U.S. Forest Service
USGS
U.S. Geological Survey
UV
ultraviolet
UV-A
ultraviolet radiation of

wavelengths from 320 to 400 nm
UV-B
ultraviolet radiation of

wavelengths from 280 to 320 nm
Vd
deposition rate, deposition

velocity (cm/s)
voc
volatile organic compound
VSD
Very Simple Dynamic (soil

acidification model)
VTSSS
Virginia Trout Stream

Sensitivity Study
WARMS
Waterfowl Acidification

Response Modeling System
WATERSN
Watershed Assessment Tool for

Evaluating Reduction Scenarios

for Nitrogen
WBB
West Bear Brook
WEBB
Water, Energy, and

Biogeochemical Budgets
WFPS
water-filled pore space
WGE
Working Group on Effects
WLS
Western Lakes Survey
June 2018
xlix
DRAFT: Do Not Cite or Quote

-------
Acronym/Abbreviation
WMO
WMP
WSA
wt %
WTA
WTP
XNOs
XO
yr
Zn
ZnO
Meaning
World Meteorological
Organization
Watershed Manipulation Project
Wadeable Stream Assessment
(survey)
percent by weight
willingness-to-accept
willingness-to-pay
nitrate halogen-X salt
halogen-X oxide
year
zinc
zinc oxide
June 2018
1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
PREFACE
Legislative Requirements for the Review of the National Ambient
Air Quality Standards
Two sections of the Clean Air Act (CAA) govern the establishment, review, and revision
of the National Ambient Air Quality Standards (NAAQS). Section 108 [42 U.S. Code
(U.S.C.) 7408] directs the Administrator to identify and list certain air pollutants and then
to issue air quality criteria for those pollutants. The Administrator is to list those air
pollutants that in their "judgment, cause or contribute to air pollution which may
reasonably be anticipated to endanger public health or welfare," "the presence of which
in the ambient air results from numerous or diverse mobile or stationary sources," and
"for which ... [the Administrator] plans to issue air quality criteria ..." [42 U.S.C.
7408(a)(1); (CAA. 1990a)1. Air quality criteria are intended to "accurately reflect the
latest scientific knowledge useful in indicating the kind and extent of all identifiable
effects on public health or welfare, which may be expected from the presence of [a]
pollutant in the ambient air ..." [42 U.S.C. 7408(b)]. Section 109 [42 U.S.C. 7409;
(CAA. 1990bVI directs the Administrator to propose and promulgate "primary" and
"secondary" NAAQS for pollutants for which air quality criteria are issued.
Section 109(b)(1) defines a primary standard as one "the attainment and maintenance of
which in the judgment of the Administrator, based on such criteria and allowing an
adequate margin of safety, are requisite to protect the public health."1 A secondary
standard, as defined in Section 109(b)(2), must "specify a level of air quality the
attainment and maintenance of which, in the judgment of the Administrator, based on
such criteria, is requisite to protect the public welfare from any known or anticipated
adverse effects associated with the presence of [the] air pollutant in the ambient air."2
In setting standards that are "requisite" to protect public health and welfare as provided in
Section 109(b), the U.S. EPA's task is to establish standards that are neither more nor less
stringent than necessary for these purposes. In so doing, the U.S. EPA may not consider
the costs of implementing the standards.3 Likewise, "[a]ttainability and technological
1	The legislative history of Section 109 indicates that a primary standard is to be set at"... the maximum permissible
ambient air level ... which will protect the health of any [sensitive] group of the population," and that for this
purpose "reference should be made to a representative sample of persons comprising the sensitive group rather than
to a single person in such a group" S. Rep. No. 91:1196, 91st Cong., 2d Sess. 10 (1970).
2	Section 302(h) of the Act [42 U.S.C. 7602(h)] provides that all language referring to effects on welfare includes,
but is not limited to, "effects on soils, water, crops, vegetation, man-made materials, animals, wildlife, weather,
visibility and climate, damage to and deterioration of property, and hazards to transportation, as well as effects on
economic values and on personal comfort and well-being ..." (CAA. 20051.
3	See generally, Whitman v. American Trucking Associations, 531 U.S. 457, 465-472, 475-476 (2001).
June 2018
li
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
feasibility are not relevant considerations in the promulgation of national ambient air
quality standards."1
Section 109(d)(1) requires that "not later than December 31, 1980, and at 5-year intervals
thereafter, the Administrator shall complete a thorough review of the criteria published
under Section 108 and the national ambient air quality standards ... and shall make such
revisions in such criteria and standards and promulgate such new standards as may be
appropriate ...Section 109(d)(2) requires that an independent scientific review
committee "shall complete a review of the criteria ... and the national primary and
secondary ambient air quality standards ... and shall recommend to the Administrator any
new ... standards and revisions of existing criteria and standards as may be
appropriate ...Since the early 1980s, this independent review function has been
performed by the Clean Air Scientific Advisory Committee (CASAC).2
Overview and History of the Reviews of the Secondary National
Ambient Air Quality Standards for Nitrogen Dioxide, Sulfur
Dioxide, and Particulate Matter
NAAQS are defined by four basic elements: indicator, averaging time, level, and form.
The indicator defines the pollutant to be measured in the ambient air for the purpose of
determining compliance with the standard. The averaging time defines the time period
over which air quality measurements are to be obtained and averaged or cumulated,
considering evidence of effects associated with various time periods of exposure. The
level of a standard defines the air quality concentration used (i.e., an ambient
concentration of the indicator pollutant) in determining whether the standard is achieved.
The form of the standard defines the air quality statistic that is compared to the level of
the standard in determining whether an area attains the standard. The Administrator
considers these four elements collectively in evaluating the protection to public health
provided by the primary NAAQS.
Nitrogen Dioxide Secondary National Ambient Air Quality
Standards
The first air quality criteria and standards for oxides of nitrogen were issued in 1971
rOJ.S. EPA. 1971). 36 FR 8186], Both the primary and secondary standards were set at
0.053 parts per million (ppm), as an annual arithmetic mean (36 FR 8186). In 1982, the
1	See American Petroleum Institute v. Costle, 665 F. 2d at 1185.
2	Lists of chartered CASAC members and of members of the CASAC Panels are available at:
http://vosemite.epa.gov/sab/sabproduct.nsfAVebCASAC/CommitteesandMembership70penDocument.
June 2018
lii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
U.S. EPA published Air Quality Criteria for Oxides of Nitrogen (U.S. EPA. 1982a).
which updated the scientific criteria upon which the initial standards were based. On
February 23, 1984, the U.S. EPA proposed to retain these standards (49 FR 6866). After
taking into account public comments, the U.S. EPA published the final decision to retain
the existing standards on June 19, 1985 (50 FR 25532).
In November 1991, the U.S. EPA initiated another review and released an updated draft
air quality criteria document (AQCD) for review and comment by CASAC and the public
(56 FR 59285). The final AQCD was released later in 1993 (U.S. EPA. 1993). Staff of
the Office of Air Quality Planning and Standards (OAQPS) prepared a draft Staff Paper
that summarized and integrated the key studies and scientific evidence contained in the
revised air quality criteria document and identified the critical elements to be considered
in the review of the NO2 NAAQS. The Staff Paper was reviewed by the CASAC and the
public in December 1994, and in September 1995, the U.S. EPA finalized the Staff Paper
(U.S. EPA. 1995b). On October 2, 1995, the Administrator announced her proposed
decision not to revise either the primary or secondary NAAQS for NO2 based on the
information available in this review (60 FR 52874; October 11, 1995). After
consideration of public comments, the Administrator made a final determination that
revisions to neither the primary nor the secondary NAAQS for NO2 were appropriate at
that time (61 FR 52852; October 8, 1996).
The most recent review of the secondary NAAQS standards for oxides of nitrogen was
performed jointly with a review of the secondary NAAQS for oxides of sulfur beginning
in 2005 (described below).
Sulfur Dioxide Secondary National Ambient Air Quality
Standards
Based on the 1969 sulfur oxides criteria document (HEW. 1969). the U.S. EPA
promulgated the initial primary and secondary NAAQS for SO2 on April 30, 1971 (36 FR
8186). The secondary standards were 0.02 ppm as an annual arithmetic mean and
0.5 ppm as a maximum 3-hour, not to be exceeded more than once per year. These
secondary standards were established on the basis of vegetation effects evidence
described in the 1970 criteria document. Based on additional data available in 1973,
revisions were made to Chapter 5 "Effects of Sulfur Oxide in the Atmosphere on
Vegetation" of the Air Quality Criteria for Sulfur Oxides (U.S. EPA. 1973). which led the
U.S. EPA to propose (38 FR 11355) and then finalize a revocation of the annual mean
secondary standard (38 FR 25678). At that time, the U.S. EPA additionally considered
welfare effects related to effects on materials, visibility, soils, and water. However, the
liii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
U.S. EPA concluded that either protection from such effects was afforded by the primary
standard or that sufficient data were not then available to develop criteria for standards
based on these effects (38 FR 25680).
In 1980, the U.S. EPA released a combined AQCD for sulfur oxides and particulate
matter for CASAC review. Following its review of a draft revised criteria document in
August 1980, the CASAC concluded that acidic deposition was atopic of extreme
scientific complexity, noting that a fundamental problem of addressing acid deposition in
a criteria document is that acidic deposition is produced by several pollutants, including
oxides of sulfur, oxides of nitrogen, and the fine particulate fraction of suspended
particles lYU.S. EPA. 1982b). pp. 125-126], Following CASAC closure on the criteria
document in December 1981, the U.S. EPA released a final AQCD (U.S. EPA. 1982b).
and the OAQPS prepared a Staff Paper that was released in November 1982 (U.S. EPA.
1982c). The issue of acidic deposition was not, however, assessed directly in the OAQPS
Staff Paper because the U.S. EPA followed the guidance given by CASAC.
In response to CASAC recommendations for a separate comprehensive discussion of
acidic deposition as part of the criteria documents, the U.S. EPA subsequently prepared
the following documents: The Acidic Deposition Phenomenon and Its Effects: Critical
Assessment Review Papers, Volumes I andII (U.S. EPA. 1984a. b) and The Acidic
Deposition Phenomenon and Its Effects: Critical Assessment Document I (Bennett ct al..
1985); 53 FR 14935-14936], Although these documents were not considered criteria
documents and had not undergone CASAC review, they represented the most
comprehensive summary of relevant scientific information completed by the U.S. EPA at
that point (58 FR21355).
At about the same time in 1980 as the CASAC recommendation for a comprehensive
assessment of acidic deposition, Congress created the National Acid Precipitation
Assessment Program (NAPAP). During the 10-year course of this program, a series of
reports were issued and a final report was issued in 1990 (NAPAP. 1991).
On April 26, 1988, the U.S. EPA proposed not to revise the existing primary and
secondary standards. This proposal regarding the secondary SO2 NAAQS was due to the
Administrator's conclusions that (1) based upon the then-current scientific understanding
of the acidic deposition problem, it would be premature and unwise to prescribe any
regulatory control program at that time and (2) when the fundamental scientific
uncertainties had been reduced through ongoing research efforts, the U.S. EPA would
draft and support an appropriate set of control measures (53 FR 14926). Subsequent to
the proposal, Congress took up consideration of acidic deposition.
liv
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
On November 15, 1990, Amendments to the CAA were passed by Congress and signed
into law by the President. In Title IV of these Amendments, Congress included a
statement of findings that had led them to take this action, including that: "(1) the
presence of acidic compounds and their precursors in the atmosphere and in deposition
from the atmosphere represents a threat to natural resources, ecosystems, materials,
visibility, and public health; (2) the problem of acid deposition is of national and
international significance; and that (3) current and future generations of Americans will
be adversely affected by delaying measures to remedy the problem..." The goal of
Title IV was to reduce emissions of SO2 by 10 million tons and oxides of nitrogen
emissions by 2 million tons from 1980 emission levels in order to achieve reductions over
broad geographic regions/areas. In envisioning that further action might be necessary in
the long term, Congress included Section 404 of the 1990 Amendments. This section
requires the U.S. EPA to conduct a study on the feasibility and effectiveness of an acid
deposition standard or standards to protect "sensitive and critically sensitive aquatic and
terrestrial resources" and at the conclusion of the study, submit a report to Congress. Five
years later, the U.S. EPA submitted to Congress its report titled Acid Deposition Standard
Feasibility Study: Report to Congress (U.S. EPA. 1995a) in fulfillment of this
requirement. The Report to Congress concluded that establishing acid deposition
standards for sulfur and nitrogen deposition might at some point in the future be
technically feasible although appropriate deposition loads for these acidifying chemicals
could not be defined with reasonable certainty at that time.
The 1990 Amendments also added new language to sections of the CAA that pertain to
the scope or application of the secondary NAAQS designed to protect the public welfare.
Section 108(g) specified that "the Administrator may assess the risks to ecosystems from
exposure to criteria air pollutants (as identified by the Administrator at the
Administrator's sole discretion)." The definition of public welfare in Section 302(h) was
expanded to state that the welfare effects identified should be protected from adverse
effects associated with criteria air pollutants ".. .whether caused by transformation,
conversion, or combination with other air pollutants."
In response to these legislative initiatives, the U.S. EPA and other federal agencies
continued research on the causes and effects of acidic deposition and related welfare
effects of SO2 and implemented an enhanced monitoring program to track progress
(58 FR 21357). In 1993, the U.S. EPA announced a decision not to revise the secondary
standard, concluding that revision to address acidic deposition and related SO2 welfare
effects was not appropriate at that time (58 FR 21351). In reaching this decision, the
U.S. EPA took into account the significant reductions in SO2 emissions, ambient SO2
concentrations, and ultimately deposition expected to result from implementation of the
Title IV program, which was expected to significantly decrease the acidification of water
lv
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
bodies and damage to forest ecosystems and to permit much of the existing damage to be
reversed with time (58 FR 21357). While recognizing that further action might be needed
to address acidic deposition in the longer term, the U.S. EPA judged it prudent to await
the results of the studies and research programs then underway, including those assessing
the comparative merits of secondary standards, acidic deposition standards, and other
approaches to control of acidic deposition and related effects, and then to determine
whether additional control measures should be adopted or recommended to Congress
(58 FR21358).
In 2000, the U.S. EPA announced receipt of two items related to acidic deposition and the
NAAQS (65 FR 48699). The first was a petition submitted to the U.S. EPA in 1999 by
representatives of seven northeastern states for the promulgation of revised secondary
NAAQS for the criteria pollutants associated with the formation of acid rain (including
NO2, SO2, and fine particulate matter [PM2 5]). The petition states that the language in
Section 302(h) of the CAA "clearly references the transformation of pollutants resulting
in the inevitable formation of sulfate and nitrate aerosols and/or their ultimate
environmental impacts as wet and dry deposition, clearly signaling Congressional intent
that the welfare damage occasioned by sulfur and nitrogen oxides be addressed through
the secondary standard provisions of Section 109 of the Act." The petition further stated
that "recent federal studies, including the NAPAP Biennial Report to Congress: An
Integrated Assessment, document the continued—and increasing—damage being
inflicted by acid deposition to the lakes and forests of New York, New England, and
other parts of our nation, demonstrating that the Title IV program had proven
insufficient." The petition also listed other adverse welfare effects associated with the
transformation of these criteria pollutants, including visibility impairment, eutrophication
of coastal estuaries, global warming, tropospheric ozone, and stratospheric ozone
depletion.
The second item was a related request from the U.S. Department of Interior (DOI) that
the U.S. EPA address many of the same adverse environmental effects associated with
the same types of air pollutants and with ozone that the DOI asserted were occurring in
national parks and wilderness areas (65 FR 48699). Included among the effects of
concern identified in the request were acidification of streams, surface waters and/or
soils, eutrophication of coastal waters, visibility impairment, and foliar injury from ozone
(65 FR 48701). The U.S. EPA requested comment on the issues raised by these requests,
stating that it would consider any relevant comments and information submitted, along
with the information provided by the petitioners and DOI, before making any decision
concerning a response to these requests for rulemaking, which if commenced would
include opportunity for public review and comment (65 FR 48701).
lvi
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Particulate Matter Secondary National Ambient Air Quality
Standards
The U.S. EPA first established NAAQS for particulate matter (PM) in 1971 (36 FR 8186,
April 30, 1971) based on the original AQCD (NAPCA. 1969). The AQCD assessed the
evidence for a variety of PM-associated welfare effects, including visibility impairment
and materials damage (e.g., soiling, corrosion). Based on evidence for such effects, the
secondary standards were set at 150 |ig/m3 for the 24-hour average not to be exceeded
more than once per year and 60 |ig/m3 for the annual geometric mean. The federal
reference method (FRM) specified for determining attainment of the original standards
was the high-volume sampler, which collects PM up to a nominal size of 25 to
45 micrometers (|im: referred to as total suspended particulates or TSP).
In October 1979 (44 FR 56730; October 2, 1979), the U.S. EPA announced the first
periodic review of the air quality criteria and NAAQS for PM. Revised primary and
secondary standards were promulgated in 1987 (52 FR 24634; July 1, 1987). In the 1987
decision, the U.S. EPA changed the indicator for particles from TSP to PMio to focus on
the subset of inhalable particles small enough to penetrate to the thoracic region of the
respiratory tract (including the tracheobronchial and alveolar regions) referred to as
thoracic particles.1 The level of the 24-hour standards (primary and secondary) was set at
150 |ig/m3. and the form was one expected exceedance per year, on average, over 3 years.
The level of the annual standards (primary and secondary) was set at 50 |ig/m3. and the
form was annual arithmetic mean averaged over 3 years.
In April 1994, the U.S. EPA announced its plans for the second periodic review of the air
quality criteria and NAAQS for PM, and in 1997, the U.S. EPA promulgated revisions to
the NAAQS (62 FR 38652, July 18, 1997). In the 1997 decision, the U.S. EPA
determined that the fine and coarse fractions of PMio should be considered separately.
This determination was based on evidence that serious health effects were associated with
short- and long-term exposures to fine particles in areas that met the existing PMio
standards. The U.S. EPA added new standards using PM2 5 as the indicator for fine
particles (with PM2 5 referring to particles with a nominal mean aerodynamic diameter
less than or equal to 2.5 |im). These new standards were as follows: (1) an annual
standard with a level of 15.0 |ig/m3 based on the 3-year average of annual arithmetic
mean PM2 5 concentrations from single or multiple community-oriented monitors and
(2) a 24-hour standard with a level of 65 |ig/m3 based on the 3-year average of the 98th
1 PMio refers to particles with a nominal mean aerodynamic diameter less than or equal to 10 |im. More specifically,
10 |im is the aerodynamic diameter for which the efficiency of particle collection is 50%. Larger particles are not
excluded altogether but are collected with substantially decreasing efficiency while smaller particles are collected
with increasing efficiency.
June 2018
lvii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
percentile of 24-hour PM2 5 concentrations at each monitor within an area. Also, the
U.S. EPA established a new reference method for the measurement of PM2 5 in the
ambient air and adopted rules for determining attainment of the new standards. To
continue to address the coarse fraction of PM10 (referred to as thoracic coarse particles or
PM10-2.5; generally including particles with a nominal mean aerodynamic diameter
greater than 2.5 |im and less than or equal to 10 |im). the U.S. EPA retained the annual
PM10 standard and revised the form of the 24-hour PM10 standard to be based on the 99th
percentile of 24-hour PM10 concentrations at each monitor in an area. The U.S. EPA
revised the secondary standards by setting them equal in all respects to the primary
standards.
Following promulgation of the 1997 PM NAAQS, petitions for review were filed by a
large number of parties, addressing a broad range of issues. In May 1999, the U.S. Court
of Appeals for the District of Columbia Circuit (D.C. Circuit) upheld the U.S. EPA's
decision to establish fine particle standards, holding that "the growing empirical evidence
demonstrating a relationship between fine particle pollution and adverse health effects
amply justifies establishment of new fine particle standards." American Trucking
Associations v. EPA, 175 F. 3d 1027, 1055-56 (D.C. Cir., 1999). The D.C. Circuit also
found "ample support" for the U.S. EPA's decision to regulate coarse particle pollution,
but vacated the 1997 PM10 standards, concluding that the U.S. EPA had not provided a
reasonable explanation justifying use of PM10 as an indicator for coarse particles (175 F.
3d at 1054-55). Pursuant to the D.C. Circuit's decision, the U.S. EPA removed the
vacated 1997 PM10 standards, and the pre-existing 1987 PM10 standards remained in
place (65 FR 80776, December 22, 2000). The D.C. Circuit also upheld the U.S. EPA's
determination not to establish more stringent secondary standards for fine particles to
address effects on visibility (175 F. 3d at 1027).
The D.C. Circuit also addressed more general issues related to the NAAQS, including
issues related to the consideration of costs in setting NAAQS and the U.S. EPA's
approach to establishing the levels of NAAQS. Regarding the cost issue, the court
reaffirmed prior rulings holding that in setting NAAQS the U.S. EPA is "not permitted to
consider the cost of implementing those standards" (Id. at 1040-41). Regarding the levels
of NAAQS, the court held that the U.S. EPA's approach to establishing the level of the
standards in 1997 (i.e., both for PM and for the ozone NAAQS promulgated on the same
day) effected "an unconstitutional delegation of legislative authority" (Id. at 1034-40).
Although the court stated that "the factors U.S. EPA uses in determining the degree of
public health concern associated with different levels of ozone and PM are reasonable," it
remanded the rule to the U.S. EPA, stating that when the U.S. EPA considers these
factors for potential nonthreshold pollutants "what U.S. EPA lacks is any determinate
criterion for drawing lines" to determine where the standards should be set.
lviii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
The D.C. Circuit's holding on the cost and constitutional issues were appealed to the U.S.
Supreme Court. In February 2001, the Supreme Court issued a unanimous decision
upholding the U.S. EPA's position on both the cost and constitutional issues. Whitman v.
American Trucking Associations, 531 U.S. 457, 464, 475-76. On the constitutional issue,
the Court held that the statutory requirement that NAAQS be "requisite" to protect public
health with an adequate margin of safety sufficiently guided the U.S. EPA's discretion,
affirming the U.S. EPA's approach of setting standards that are neither more nor less
stringent than necessary.1
In October 1997, the U.S. EPA published its plans for the third periodic review of the air
quality criteria and NAAQS for PM (62 FR 55201; October 23, 1997). On September 21,
2006, the U.S. EPA announced its final decisions to revise the primary and secondary
NAAQS for PM to provide increased protection of public health and welfare,
respectively (71 FR 61144; October 17, 2006). With regard to the primary and secondary
standards for fine particles, the U.S. EPA revised the level of the 24-hour PM2 5 standards
to 35 |ig/m3. retained the level of the annual PM2 5 standards at 15.0 |ig/m3. and revised
the form of the annual PM2 5 standards by narrowing the constraints on the optional use of
spatial averaging. With regard to the primary and secondary standards for PM10, the
U.S. EPA retained the 24-hour standards, with levels at 150 |ig/m3. and revoked the
annual standards.2 The Administrator judged that the available evidence generally did not
suggest a link between long-term exposure to existing ambient levels of coarse particles
and health or welfare effects. In addition, a new reference method was added for the
measurement of PM10-2.5 in the ambient air, in order to provide a basis for approving
federal equivalent methods (FEMs) and to promote the gathering of scientific data to
support future reviews of the PM NAAQS.
Several parties filed petitions for review following promulgation of the revised PM
NAAQS in 2006. These petitions addressed the following issues: (1) selecting the level of
the primary annual PM2 5 standard; (2) retaining PM10 as the indicator of a standard for
1	The Supreme Court remanded the case to the Court of Appeals for resolution of any remaining issues that had not
been addressed in that court's earlier rulings (Id. at 475-76). In a March 2002 decision, the Court of Appeals
rejected all remaining challenges to the standards, holding that the U.S. EPA's PM2 5 standards were reasonably
supported by the administrative record and were not "arbitrary and capricious" American Trucking Associations v.
EPA, 283 F. 3d 355, 369-72 (D.C. Cir. 2002).
2	In the 2006 proposal, the U.S. EPA proposed to revise the 24-hour PM10 standard in part by establishing a new
PM10-2.5 indicator for thoracic coarse particles (i.e., particles generally between 2.5 and 10 |im in diameter). The
U.S. EPA proposed to include any ambient mix of PM10-2.5 that was dominated by resuspended dust from
high-density traffic on paved roads and by PM from industrial and construction sources. The U.S. EPA proposed to
exclude any ambient mix of PMi 0-2.5 that was dominated by rural windblown dust and soils and by PM generated
from agricultural and mining sources. In the final decision, the existing PM10 standard was retained, in part due to an
"inability.. .to effectively and precisely identify which ambient mixes are included in the [PMi0-2.5] indicator and
which are not" (71 FR 61197; October 17, 2006).
June 2018
lix
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
thoracic coarse particles, retaining the level and form of the 24-hour PMio standard, and
revoking the PMio annual standard; and (3) setting the secondary PM2 5 standards
identical to the primary standards. On February 24, 2009, the U.S. Court of Appeals for
the District of Columbia Circuit issued its opinion in the case American Farm Bureau
Federation v. EPA, 559 F. 3d 512 (D.C. Cir. 2009). The court remanded the primary
annual PM2 5 NAAQS to the U.S. EPA because the U.S. EPA failed to adequately explain
why the standards provided the requisite protection from both short- and long-term
exposures to fine particles, including protection for at-risk populations (American Farm
Bureau Federation v. EPA, 559 F. 3d 512, 520-27; D.C. Cir. 2009). With regard to the
standards for PMio, the court upheld the U.S. EPA's decisions to retain the 24-hour PMio
standard to provide protection from thoracic coarse particle exposures and to revoke the
annual PMio standard (American Farm Bureau Federation, 559 F. 2d at 533-38). With
regard to the secondary PM2 5 standards, the court remanded the standards to the
U.S. EPA because the Agency failed to adequately explain why setting the secondary PM
standards identical to the primary standards provided the required protection for public
welfare, including protection from visibility impairment (American Farm Bureau
Federation, 559 F. 2d at 528-32). The U.S. EPA responded to the court's remands as
part of the next review of the PM NAAQS, which was initiated in 2007.
In June 2007, the U.S. EPA initiated the fourth periodic review of the air quality criteria
and the PM NAAQS by issuing a call for information in the Federal Register (72 FR
35462; June 28, 2007). In December 2012, the U.S. EPA announced its final decisions
with regard to the secondary PM standards, the U.S. EPA retained the 24-hour and annual
PM2 5 standards and the 24-hour PMio standard to address visibility and nonvisibility
welfare effects. On judicial review, the revised standards were upheld in all respects
(NAMv. EPA, 750 F.3d 921; D.C. Cir. 2014).
Most Recent Combined Review of the Oxides of Nitrogen
and Oxides of Sulfur National Ambient Air Quality
Standards
In 2005, the U.S. EPA initiated a joint review of the air quality criteria for oxides of
nitrogen and sulfur and the secondary NAAQS for NO2 and SO2. In so doing, the
U.S. EPA assessed the scientific information, associated risks, and standards relevant to
protecting the public welfare from adverse effects associated jointly with oxides of
nitrogen and sulfur. Although the U.S. EPA has historically adopted separate secondary
standards for oxides of nitrogen and oxides of sulfur, the U.S. EPA conducted a joint
review of these standards because oxides of nitrogen and sulfur and their associated
transformation products are linked from an atmospheric chemistry perspective, as well as
lx
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
from an environmental effects perspective. The joint review was also responsive to the
National Research Council (NRC) recommendation for the U.S. EPA to consider
multiple pollutants, as appropriate, in forming the scientific basis for the NAAQS (NRC.
2004V
The review was initiated on December 13, 2005 with a call for information (70 FR
73236) for the development of a revised ISA. A draft Integrated Review Plan (IRP) was
released in October 2007, reviewed by CASAC; the final IRP was released in December
2007 (U.S. EPA. 2007). The first and second drafts of the ISA were released in
December 2007 and August 2008 (73 FR 10243), respectively, for CASAC and public
review. The final ISA (U.S. EPA. 2008a) was released in December 2008 (73 FR 75716).
Based on the scientific information in the ISA, the U.S. EPA developed a Risk and
Exposure Assessment (REA) to further assess the national impact of the effects
documented in the ISA. The Draft Scope and Methods Plan for Risk/Exposure
Assessment: Secondary NAAQS Review for Oxides of Nitrogen and Oxides of Sulfur
outlining the scope and design of the future REA was released in March 2008 (73 FR
10243). A first and second draft of the REA were released (August 2008 and June 2009)
for CASAC review and public comment. The final REA (U.S. EPA. 2009c) was released
in September 2009. Drawing on the information in the final REA and ISA, a first draft,
second draft, and final Policy Assessment (PA) were released in March 2010, September
2010, and January 2011, respectively (U.S. EPA. 2011a).
On August 1, 2011, based on consideration of the scientific information and quantitative
assessments, the U.S. EPA published a proposal to (1) retain the existing NO2 and SO2
secondary standards, (2) add secondary standards identical to the NO2 and SO2 primary
1-hour standards, and (3) not set a new multipollutant secondary standard in this review.
After consideration of public comments on the proposed standards and on design of a
new field pilot program to gather and analyze additional relevant data, the Administrator
signed a final decision in this rulemaking on March 20, 2012. The Administrator's
decision was that, while the current secondary standards were inadequate to protect
against adverse effects from deposition of oxides of nitrogen and sulfur, it was not
appropriate under Section 109(b) to set any new secondary standards at this time due to
the limitations in the available data and uncertainty as to the amount of protection the
metric developed in the review would provide against acidification effects across the
country (77 FR 20281). In addition, the Administrator decided that it was appropriate to
retain the current NO2 and SO2 secondary standards to address direct effects of gaseous
NO2 and SO2 on vegetation. Thus, taken together, the Administrator decided to retain and
not revise the current NO2 and SO2 secondary standards: an NO2 standard set at a level of
lxi
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
0.053 ppm as an annual arithmetic average, and an SO2 standard set at a level of 0.5 ppm
as a 3-hour average, not to be exceeded more than once per year (77 FR 20281).
The U.S. EPA's decision to not set a secondary NAAQS for oxides of nitrogen and sulfur
even though the Administrator had concluded that the existing standards are not adequate
to protect against the adverse impacts of aquatic acidification on sensitive ecosystems
was challenged by the Center for Biological Diversity and other environmental groups.
The petitioners argued that having decided that the existing standards were not adequate
to protect against adverse public welfare effects such as damage to sensitive ecosystems,
the Administrator was required to identify the requisite level of protection for the public
welfare and to issue a NAAQS to achieve and maintain that level of protection. The D.C.
Circuit disagreed, finding that the U.S. EPA acted appropriately in not setting a
secondary standard given the U.S. EPA's conclusions that "the available information was
insufficient to permit a reasoned judgment about whether any proposed standard would
be 'requisite to protect the public welfare ..."' (Center for Biological Diversity, etal. v.
EPA, 749 F.3d 1079, 1087; 2014). In reaching this decision, the court noted that the
U.S. EPA had "explained in great detail" the profound uncertainties associated with
setting a secondary NAAQS to protect against aquatic acidification.
lxii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
EXECUTIVE SUMMARY
Purpose and Scope of the Integrated Science Assessment
This Integrated Science Assessment (ISA) for Oxides of Nitrogen, Oxides of Sulfur, and
Particulate Matter—Ecological Criteria is a comprehensive evaluation and synthesis of
the most policy-relevant science aimed at characterizing the ecological effects caused by
these criteria pollutants.1 These criteria pollutants are reviewed here together because
they all contribute to nitrogen (N) and sulfur (S) deposition, which causes substantial
ecological effects. In this document, the term "oxides of nitrogen" refer to total oxidized
N (NOy), including nitric oxide (NO) and nitrogen dioxide (NO2) and all other oxidized
N containing compounds formed from NO and NO2.2 The term "oxides of sulfur"
includes gaseous sulfur oxides (e.g., sulfur dioxide [SO2], sulfur monoxide [SO], disulfur
monoxide [S2O], and sulfur trioxide [SO3]) as well as particulate species, such as
ammonium sulfate [(NH4)2S04 (U.S. EPA. 2011aYI. Particulate species include oxidized
sulfur species like sulfites (SO32 ) and sulfates (SO42 ). but among these two species
usually only sulfates make a major contribution to particulate mass. Throughout this
document SOx is defined as the sum of SO2 and particulate sulfate (SO42 ). which
represent virtually all of the oxidized sulfur mass in the atmosphere.3 Particulate matter
(PM) is composed of some or all of the following components: nitrate (NO3 ). SO42 .
ammonium (NH4+), metals, minerals (dust), and organic and elemental carbon.
This ISA serves as the scientific foundation for the review of the ecological effects
associated with the secondary (welfare-based) National Ambient Air Quality Standards
(NAAQS) for oxides of nitrogen, oxides of sulfur, and particulate matter. The health
effects of these criteria pollutants are considered in separate assessments as part of the
review of the primary (health-based) NAAQS for oxides of nitrogen (U.S. EPA. 2016f).
oxides of sulfur (U.S. EPA. 2016e). and particulate matter (U.S. EPA. 2009a).4 The
1	The general process for developing an ISA, including the framework for evaluating weight of evidence and
drawing scientific conclusions and causal judgments, is described in a companion document, Preamble to the
Integrated Science Assessments (U.S. EPA. 2015e'). www.epa.gov/isa.
2	This ISA reserves the abbreviation NOx strictly as the sum of NO and NO2—consistent with its use in the
atmospheric science community—and uses the term "oxides of nitrogen" to refer to the broader list of oxidized
nitrogen species. Oxides of nitrogen refers to NOy as the total oxidized nitrogen in both gaseous and particulate
forms. The major gaseous and particulate constituents of NOy include nitric oxide (NO), nitrogen dioxide (NO2),
nitric acid (HNO3), peroxyacetyl nitrate (PAN), nitrous acid (HONO), organic nitrates, and particulate nitrate (NO3).
This ISA uses the definitions adopted by the atmospheric sciences community.
3	The same definition of SOx used in the 2011 NOx SOx Policy Assessment (U.S. EPA. 201 la").
4	In this ISA, the blue electronic links can be used to navigate to cited materials as well as appendices, chapters,
sections, tables, figures, and studies from this ISA.
June 2018
lxiii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Clean Air Act definition of welfare effects includes, but is not limited to, effects on soils,
water, wildlife, vegetation, visibility, weather, and climate, as well as effects on
man-made materials, economic values, and personal comfort and well-being.
The current secondary NAAQS for oxides of nitrogen and oxides of sulfur were set to
protect against direct damage to vegetation by NO2 or SO2. The secondary NAAQS for
NO2 is identical to the primary standard set in 1971: an annual average not to exceed
0.053 ppm nitrogen dioxide. The secondary NAAQS for SO2, set in 1973, is a 3-hour
average of 0.5 ppm SO2, not to be exceeded more than once per year. The current
secondary standards for PM are intended to address PM-related visibility and
nonvisibility welfare effects. These standards are a 3-year annual mean PM2 5
concentration of 15 |ig/m3. with the 24-hour average PM2 5 and PM10 set at concentrations
of 35 |ig/m3 and 150 |ig/m3. respectively.
This ISA updates the 2008 ISA for Oxides of Nitrogen and Oxides of Sulfur—Ecological
Criteria [hereafter referred to as 2008 ISA (U.S. EPA. 2008aVI. as well as the ecological
portion of the 2009 ISA for Particulate Matter (U.S. EPA. 2009a) with studies and
reports published from January 2008 through May 2017. Some studies published more
recently than May 2017 are also included based on expert judgement. The U.S. EPA
conducted in-depth searches to identify peer-reviewed literature on relevant topics.
Subject-area experts and the public were also able to recommend studies and reports
during a kick-off workshop held at the U.S. EPA in March 2014 for oxides of nitrogen
and oxides of sulfur, and June 2016 for PM. The Clean Air Scientific Advisory
Committee (CASAC) recommended literature during the review of the first draft. To
fully describe the state of available science, the U.S. EPA also carried over the most
relevant studies from previous assessments to include in this ISA.
This ISA determines whether NOy, SOx, and PM concentrations in the air or depositing
from the air cause ecological effects. The ecological effects of deposition are grouped
into three main categories: (1) acidification (caused by NOy, SOx, and particulate forms
of N and S), (2) N enrichment/N driven eutrophication (caused by NOy and particulate
forms ofN), and (3) S enrichment (caused by SOx and particulate forms of S). Ecological
effects are further subdivided into terrestrial, wetland, freshwater, and
estuarine/near-coastal ecosystems. These ecosystems and effects are linked by the
connectivity of terrestrial and aquatic habitats through biogeochemical pathways of N
and S.
A schematic of the document organization is given by Figure ES-1. Chapter 1 is the
integrative chapter that brings together key information on specific subject matter found
in the Appendices. Appendix 1 is an introduction to the purpose and organization of
Appendix 2-Appendix 16. Appendix 2 characterizes the sources and atmospheric
lxiv
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
processes involving NOy, SOx, and PM, as well as trends in ambient concentrations and
deposition. Appendix 3 describes direct effects of gas-phase NOy and SOx on plants and
lichens. Appendix 4 describes N and S deposition effects on terrestrial biogeochemistry,
and Appendix 5 and Appendix 6 describe the biological effects of terrestrial acidification
and terrestrial N enrichment, respectively. Appendix 7 describes N and S deposition
effects on aquatic biogeochemistry. Appendix 8 through Appendix 10 respectively
characterize the biological effects of freshwater acidification, freshwater N enrichment,
and marine eutrophication. Appendix 11 describes the effects of N deposition on
wetlands. Appendix 12 describes the wetland and freshwater effects of S enrichment.
Appendix 13 discusses the climate modification of ecosystem response to N and S
deposition, and Appendix 14 presents information on N and S deposition effects on
ecosystem services. Information on the ecological effects of forms of PM beyond those
related to N or S deposition, is presented in Appendix 15 [the nonecological welfare
effects associated with PM, such as visibility, climate, and material effects, are
considered as part of a separate review of PM (81 FR 87933, December 6, 2016)].
Appendix 16 includes six locations in the U.S. selected as case study areas that are
candidates for additional analysis of risk and exposure. These candidate sites were
selected because they have abundant data on ecological effects.
lxv
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
Integrative
Chapter
(Chapter 1)
Exposure
1 Biological Effects 1

Ambient Air Concentrations


so2, no2, no, pan, hno3


(Appendix 2)


(Appendix 3)


Direct exposure


Plant foliar and lichen Injury

Deposition of
N and S
(Appendix 2)
Soil and aquatic biogeochemical pathways of
acidification (NOY+ NHX +SOx)
N enrichment/eutrophication (NOY +NHX)
S-nutrient (SOx)
Terrestrial Ecosystems
Direct to organism/deposition
Direct to soil, effects on soil biogeochemistry (Appendix 4)
Wetland Ecosystems
Directto soil and surface water, runoff from soil
Wetland biogeochemistry (Appendices 11&12)
Freshwater Ecosystems
Directto surface water, runoff from soil, effects on
freshwater biogeochemistry (Appendix 7)
Estuaries Ecosystems
Directto water, transport from watershed runoff, effects on
biogeochemistry along the freshwater to ocean continuum
(Appendix 7)
Biological effects of
acidification (NOY+ NHX +SOx)
N enrichment/eutrophication (NOy +NHX)
S-nutrient (S0X)
Terrestrial Ecosystems
Acidification (Appendix 5)
N enrichment/eutrophication (Appendix 6)
Wetlands Ecosystems
N enrichment/eutrophication (Appendix 11}
	S-nutrient (Appendix 12)	
Freshwater Ecosystems
Acidification (Appendix 8)
N enrichment/eutrophication (Appendix 9)
	S-nutrient (Appendix 12)	
Estuarine Ecosystems
N-nutrient/ eutrophication (Appendix 10)
N-enhanced ocean acidification (Appendix 10)
Climate Modification of Ecosystem
Response to N and S
(Appendix 13)
Ecosystem
Services
(Appendix 14)
Other Ecological
Effects of PM
(Appendix 15)
Case
Studies
(Appendix 16)
HN03 = nitric acid; N = nitrogen; NHX = reduced nitrogen; NO = nitric oxide; N02 = nitrogen dioxide; NOx = NO + N02;
PAN = peroxyacetyi nitrate; PM = particulate matter; S = sulfur; S02 = sulfur dioxide; SOx = sulfur oxides.
Figure ES-1 Roadmap of the Integrated Science Assessment linking
atmospheric concentrations and deposition, soil and aquatic
biogeochemistry, and biological effects.
Emissions, Ambient Air Concentrations, and Deposition
The atmospheric chemistry from emission to deposition discussed in this ISA1 is for the
criteria pollutants NOy, SOx, and PM. NOy and SOx cause ecological effects in the gas
phase and/or by N and S deposition to an ecosystem surface. Particulate matter (PM)
effects discussed in this document are also mainly focused on N and S containing species,
which together usually make up most of the PM? s mass in most areas of the U.S. NHx
(NHx = NH3 + NH4+) includes both NH3 and the PM component NH4+. NH3 is estimated
to account for 19-63% of total observed inorganic N deposition; therefore, the
contribution of NOy, PM component NH4 . and NH3 to N deposition is discussed in this
ISA.
Both gaseous and particulate forms of NOy, SOx, and NHx contribute to atmospheric
deposition. The major components of PM in the U.S. are NO-- . SO r . NH4+, organic
1 The term concentration is used throughout the ISA to denote either a mass per unit volume or a mixing ratio. The
use of concentration to denote abundance expressed as mixing ratio is firmly entrenched in the literature; therefore,
it is retained here, despite being technically incorrect.
June 2018
lxvi
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
carbon, and elemental carbon. Of these, NO3 , SO42 . and NH4+ usually have a strong
influence on acid deposition, and NO3 and NFU+, and in some cases organic N (organic
nitrates and reduced organic N), make a substantial contribution to N deposition.
The sources and precursors to gaseous and particulate forms of NOy, SOx, and NHx vary.
The main contributors to acidifying precipitation are formed from precursor emissions of
the gases SO2 and NOx (NO + NO2) (Appendix 2.2). Electricity generating units (EGUs)
are the source of most gaseous emissions of SO2. Notably, SO2 emissions from EGUs
have been decreasing. NOx emissions have a wider distribution of sources, with
substantial contributions from highway and off-highway vehicles, lightning, and EGUs.
Agriculture (fertilizer application and animal waste) is the main source of NH3. Primary
PM2 5 and PM10 emissions are dominated by dust and fires, but much of the PM2 5 mass in
the U.S. is produced by reaction of gas-phase precursors to form secondary PM2 5, and the
majority of the mass is often due to N and S species produced by secondary PM2 5
formation. The particulates NH4+, NO3 . and SO42 are primarily derived from the
gaseous precursors NH3, NOx, and SO2 (Appendix 2.3). In the Eastern U.S., NO3 and
S042 make up an even greater portion of PM2 5 mass in areas where PM2 5 mass is the
highest. Formation of particulate N and S is described in the 2009 ISA for Particulate
Matter (U.S. EPA. 2009a'). An understanding of the sources, chemistry, and atmospheric
processes for these gas-phase and PM species is necessary to understand acidifying and N
deposition.
Since the passage of the Clean Air Act Amendments in 1990, emissions of NOx
(NO + NO2) and SO2 have declined dramatically. Total emissions of SO2 decreased by
72% from 1990 to 2011. Emissions ofNOx in the U.S. from highway vehicles and fuel
combustion declined 49% between 1990 and 2013, while nationwide annual average NO2
concentrations decreased by 48% from 1990 to 2012 (U.S. EPA. 2016f). This has in turn
led to decreases in PM2 5 concentrations because of a decline in SO42 formed in the
atmosphere.
Averaged across the U.S., deposition of total N (oxidized + reduced N, in kg N/ha/yr) has
not changed over the past 25 years (Appendix 2.1). Although NOx emissions have
declined in the continental U.S. (CONUS), emissions of NH3 have increased in many
areas. There is large spatial variability in N deposition over the CONUS (Chapter 1,
Figure 1-5). According to National Atmospheric Deposition Program Total Deposition
Committee (TDEP) estimates for 2011-2013 (Appendix 2.7). at least one-third of the
CONUS is estimated to receive at least 10 kg N/ha/yr dry + wet deposition, with some
areas receiving more than 15 kg N/ha/yr. Estimates are likely too low for areas receiving
at least 10 kg N/ha/yr of deposition and for the overall amount of N deposited because
lxvii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
reduced organic N species are not routinely monitored or considered in air quality models
such as the Community Multiscale Air Quality Modeling System (CMAQ).
For S, wet deposition tends to dominate over dry deposition in large areas of the CONUS.
Dry deposition of particulate SOr is only a minor source of S, but there is considerably
more uncertainty associated with dry deposition to ecosystems than wet deposition.
Anthropogenic emissions of S and subsequent deposition have declined markedly since
the 1990s, with the most pronounced declines in the eastern U.S. Currently, the highest
values of total (wet + dry) SOx deposition in the U.S. are in parts of the Ohio Valley
region, and range between 15 and 20 kg S/ha/yr.
Both N and S deposition contribute to acidification of ecosystems. The acidity of
rainwater has decreased, as indicated by the increase of rainwater pH across the U.S.
since 1990, coincident with decreases in the wet deposition of nitrate and sulfate.
However, widespread areas are still affected by acidifying precipitation, mainly in the
eastern U.S. (see Appendix 2.7). Total acidifying deposition (wet + dry N + S, expressed
as H+ equivalents) fluxes for 2011 to 2013 ranged from a few hundred H+
equivalents/ha/yr over much of the western U.S. to over 1,500 H+ equivalents/ha/yr in a
broad swath encompassing the Midwest and the Mid-Atlantic regions, and in other
isolated hotspots surrounding areas of concentrated industrial or agricultural activity
(Chapter 1, Figure 1-6).
Ecological Effects
In this ISA, information on ecological effects from controlled exposure, field addition,
ambient deposition, and toxicological studies, among others, are integrated to form
conclusions about the causal nature of relationships between NOy, SOx, and PM and
ecological effects. Studies on the ecological effects are considered in relation to a range
of ambient concentration and deposition loads that are within two orders of magnitude of
current conditions [Preamble (U.S. EPA. 2015e). Section 5c]. A consistent and
transparent framework [Preamble (U.S. EPA. 20156). Table II] is applied to classify the
ecological effect evidence according to a five-level hierarchy:
1.	Causal relationship
2.	Likely to be a causal relationship
3.	Suggestive of, but not sufficient to infer, a causal relationship
4.	Inadequate to infer a causal relationship
5.	Not likely to be a causal relationship
The conclusions presented in Table ES-1 are informed by recent findings integrated with
information from the 2008 ISA (U.S. EPA. 2008a). Important considerations include
June 2018
lxviii
DRAFT: Do Not Cite or Quote

-------
1	judgments of error and uncertainty, as well as the coherence of findings integrated across
2	studies of underlying geochemical and biological mechanisms. There are 20 causality
3	statements in this ISA (Table ES-1). Fourteen are causal relationships repeated from the
4	2008 ISA or modified from the 2008 ISA to include specific endpoints. For these
5	causality statements, new research strengthens the evidence base and is consistent with
6	the 2008 ISA. There is one likely causal relationship repeated from the 2009 ISA for
7	Particulate Matter. Five are new endpoint categories not evaluated in the 2008 ISA: three
8	with causal relationships, one with a likely causal relationship, and one suggestive of a
9	causal relationship.
Table ES-1 Causal determinations for relationships between criteria pollutants
and ecological effects from the 2008 NOx/SOx ISA or the 2009 ISA for
particulate matter (PM), for other effects of PM, and the current draft
Integrated Science Assessment.

Causal Determination

Effect Category
2008 NOx/SOx ISA
Current
Draft ISA
Gas-phase direct phototoxic effects
Gas-phase SO2 and injury to vegetation
Appendix 3.5.1
Causal relationship
Causal
relationship
Gas-phase NO, NO2, and PAN and injury to
vegetation
Causal relationship
Causal
relationship
Appendix 3.5.2


Gas-phase HNO3 and injury to vegetation3
Appendix 3.5.3
Causal relationship
Causal
relationship
N and acidifying deposition to terrestrial ecosystems
N and S deposition and alteration of soil
biogeochemistry in terrestrial ecosystems'5
Causal relationship
Causal
relationship
Appendix 4.1


N deposition and the alteration of the physiology
and growth of terrestrial organisms and the
productivity of terrestrial ecosystems0
Not included
Causal
relationship
Appendix 6.6.1


N deposition and the alteration of species
richness, community composition, and biodiversity
in terrestrial ecosystems0
Causal relationship
Causal
relationship
Appendix 6.6.2


June 2018
lxix
DRAFT: Do Not Cite or Quote

-------
Table ES-1 (Continued): Causal determinations for relationships between criteria
pollutants and ecological effects from the 2008 NOx/SOx
ISA or the 2009 ISA for particulate matter (PM), for other
effects of PM, and the current draft Integrated Science
Assessment.
Causal Determination
Effect Category
2008 NOx/SOx ISA
Current
Draft ISA
Acidifying N and S deposition and the alteration of
the physiology and growth of terrestrial organisms
and the productivity of terrestrial ecosystemsd
Appendix 5.7.1
Not included
Causal
relationship
Acidifying N and S deposition and the alteration of
species richness, community composition, and
biodiversity in terrestrial ecosystemsd
Appendix 5.7.2
Causal relationship
Causal
relationship
N and acidifying deposition to freshwater ecosystems
N and S deposition and alteration of freshwater
biogeochemistrye
Appendix 7.1.7
Causal relationship
Causal
relationship
Acidifying N and S deposition and changes in
biota, including physiological impairment and
alteration of species richness, community
composition, and biodiversity in freshwater
ecosystems'
Appendix 8.6
Causal relationship
Causal
relationship
N deposition and changes in biota, including
altered growth and productivity, species richness,
community composition, and biodiversity due to N
enrichment in freshwater ecosystems9
Appendix 9.6
Causal relationship
Causal
relationship
N deposition to estuarine ecosystems
N deposition and alteration of biogeochemistry in
estuarine and near-coastal marine systems
Appendix 7.2.10
Causal relationship
Causal
relationship
N deposition and increased nutrient-enhanced
coastal acidification
Appendix 7.2.10
Not included
Likely to be
a causal
relationship
N deposition and changes in biota, including	Causal relationship	Causal
altered growth, total primary production, total algal	relationship
community biomass, species richness, community
composition, and biodiversity due to N enrichment
in estuarine environments11
Appendix 10.7.1
June 2018
lxx
DRAFT: Do Not Cite or Quote

-------
Table ES-1 (Continued): Causal determinations for relationships between criteria
pollutants and ecological effects from the 2008 NOx/SOx
ISA or the 2009 ISA for particulate matter (PM), for other
effects of PM, and the current draft Integrated Science
Assessment.
Causal Determination
Effect Category
2008 NOx/SOx ISA
Current
Draft ISA
N deposition and changes in biota, including
altered physiology, species richness, community
composition, and biodiversity due to
nutrient-enhanced coastal acidification
ADDendix 10.7.2
Not included
Suggestive
of, but not
sufficient to
infer, a
causal
relationship
N deposition to wetland ecosystems
N deposition and the alteration of biogeochemical
cycling in wetlands
ADDendix 11.10
Causal relationship
Causal
relationship
N deposition and the alteration of growth and
productivity, species physiology, species richness,
community composition, and biodiversity in
wetlands
ADDendix 11.10
Causal relationship
Causal
relationship
S deposition to wetland and freshwater ecosystems
S deposition and the alteration of mercury
methylation in surface water, sediment, and soils
in wetland and freshwater ecosystems'
Appendix 12.7
Causal relationship
Causal
relationship
S deposition and changes in biota due to sulfide
phytotoxicity, including alteration of growth and
productivity, species physiology, species richness,
community composition, and biodiversity in
wetland and freshwater ecosystems
Appendix 12.7
Not included
Causal
relationship
June 2018
lxxi
DRAFT: Do Not Cite or Quote

-------
Table ES-1 (Continued): Causal determinations for relationships between criteria
pollutants and ecological effects from the 2008 NOx/SOx
ISA or the 2009 ISA for particulate matter (PM), for other
effects of PM, and the current draft Integrated Science
Assessment.

Causal Determination

Effect Category
2008 NOx/SOx ISA
Current
Draft ISA

2009 PM ISA
Current
Draft ISA
Other ecological effects of PM
PM and a variety of effects on individual
organisms and ecosystems
Appendix 15.7
Likely to be a causal relationship
Likely to be
a causal
relationship
C = carbon; Hg = mercury; HN03 = nitric acid; ISA = Integrated Science Assessment; N = nitrogen; NO = nitric oxide;
N02 = nitrogen dioxide; PAN = peroxyacetyl nitrate; S = sulfur; S02 = sulfur dioxide.
aThe 2008 ISA causality statements for gas-phase HN03 was phrased as, "changes in vegetation."
bThe 2008 ISA included two causality statements for terrestrial biogeochemistry which were phrased as, "relationship between
acidifying deposition and changes in biogeochemistry" and "relationship between N deposition and the alteration of
biogeochemical cycling of N."
°The 2008 ISA causality statement for biological effects of N enrichment in terrestrial ecosystems was phrased as, "relationship
between N deposition and the alteration of species richness, species composition, and biodiversity."
dThe 2008 ISA causality statement for biological effects of acidifying deposition in terrestrial ecosystems was phrased as,
"relationship between acidifying deposition and changes in terrestrial biota."
eThe 2008 ISA included three causality statements for freshwater biogeochemistry phrased as, "relationship between acidifying
deposition and changes in biogeochemistry related to aquatic ecosystems," "relationship between N deposition and the alteration
of biogeochemical cycling of N," and "relationship between N deposition and the alteration of biogeochemical cycling of C."
'The 2008 ISA causality statement for biological effects of acidifying deposition in freshwater ecosystems was phrased as,
"relationship between acidifying deposition and changes in aquatic biota."
9The 2008 ISA causality statement for biological effects of N deposition in freshwater ecosystems was phrased as, "relationship
between N deposition and the alteration of species richness, species composition, and biodiversity in freshwater aquatic
ecosystems."
The 2008 ISA causality statement for biological effects of N deposition to estuaries was phrased as, "relationship between N
deposition and the alteration of species richness, species composition, and biodiversity in estuarine ecosystems."
'The 2008 ISA causality statement for biological effects of S deposition effects on ecosystems was phrased as, "relationship
between S deposition and increased methylation of Hg, in aquatic environments where the value of other factors is within
adequate range for methylation."
1
2	Figure ES-2 presents a visualization of the causality statements integrated to a single
3	diagram. There is not a one-to-one correspondence between the number of causality
4	statements, of which there are 20, and the cells indicated to have causal relationships in
5	the diagram because some causal statements include effects across more than one level of
6	biological organization. The main findings are that gaseous NOx and SOx cause
7	phytotoxic effects, while N and S deposition cause alteration of (1) biogeochemical
8	components of soil and water chemistry and (2) multiple levels of biological organization
9	ranging from physiological processes to shifts in biodiversity and ecological function.
June 2018
lxxii
DRAFT: Do Not Cite or Quote

-------
2018 NOx SOx PM Integrated Science Assessment for ECOLOGICAL EFFECTS*
Indicator
Gases Ł
Nitrogen deposition
Sulfur deposition
Nitrogen and Sulfur
deposition
UJ
o
*	A causal relationship is likely to exist between deposition of PM and a variety of effects on individual organisms and ecosystems, based
on information from the previous review and limited new findings in this review
*	Includes: NO, N02, HN03. S02, and PAN
Figure ES-2 Causal relationships between the criteria pollutants and ecological effects.
Eulrophi cation
driven
acidification
Direct
phytotoxic
Class of Pollutant Effect
N-ewchment/Euirophicatioo
Sulfide Towcity
Acidification
Mefcury Methyiation
Fresh
water
Ecosystem
Terrestnal
Terrestrial
Wetland
Estuary
Wetland
Terrestnal
Wetland
Productivity
Biodiversity
Growth rate
Physiological
alerataon. stress
or injury
Soil or sediment
chemistry
Surface water
chemistry
June 2018
lxxiii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Direct Phytotoxic Effects of Gas-Phase Oxides of Nitrogen
(NOy) and Oxides of Sulfur (SOx)
The current NO2 and SO2 secondary NAAQS are set to protect against direct damage to
vegetation by exposure to gas-phase oxides of nitrogen and oxides of sulfur. Research
continues to support causal relationships between SO2, NO2, NO, peroxyacetyl nitrate
(PAN), HNO3, and injury to vegetation (Chapter 1, Table 1-1. Section L4; Appendix 3).
but research that tests plant response to the lower exposure levels representative of
current atmospheric NOy and SOx concentrations is limited. Therefore, little evidence is
available to inform whether current monitored concentrations of gas-phase NOy and SOx
are high enough to injure vegetation.
Ecological Effects of Nitrogen and Sulfur Deposition
It is clear from the body of knowledge that NOy, SOx, and PM contribute to total N and S
deposition. In turn, N and S deposition alter ecosystem biogeochemistry as well as
organismal physiology, resulting in harmful declines in biodiversity in terrestrial,
freshwater, wetland, and estuarine ecosystems in the U.S. Decreases in biodiversity mean
that some species become relatively less abundant and may be locally extirpated. In
addition to the loss of unique living species, the decline in total biodiversity can be
harmful because biodiversity is an important determinant of the stability of ecosystems
and the ability of ecosystems to provide services to humanity (see more on biodiversity in
Chapter 1, Section 1.2.2.4).
Acidification of Terrestrial and Freshwater Ecosystems
Several decades of research have documented that N and S deposition cause freshwater
and terrestrial ecosystem acidification in the U.S. New evidence strengthens the causal
relationships for ecosystem acidification determined in the 2008 ISA (Chapter 1,
Table 1-1).
Many of the terrestrial and freshwater ecosystems most sensitive to acidification in the
U.S. are found in the Northeast and the Southeast. In the West, freshwater and terrestrial
ecosystems acidified from deposition are currently limited in extent and occur mostly in
high-elevation sites. Watershed sensitivity to acid inputs depends on characteristics such
as underlying geology (Appendix 4 and Appendix 7) and the sensitivity of species in the
local biological community (Appendix 5 and Appendix 8). Regional heterogeneity of
deposition levels that cause ecological effects are in part due to historic exposure and
climate. The effects of acidifying deposition have been studied for several decades in the
lxxiv
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Northeast and the southern Appalachian Mountains. Although S deposition is decreasing,
recent studies suggest biological recovery of previously acidified ecosystems is limited
(Chapter 1, Section 1.12).
Acidified aquatic habitats have lower species richness of fishes, macroinvertebrates, and
phytoplankton. The effects of acidifying deposition on aquatic ecosystems also include
physiological impairment or mortality of sensitive species and shifts in biodiversity of
both flora and fauna. Organisms at all trophic levels are affected by acidification, with
clear linkages to chemical indicators for effects on algae, zooplankton, benthic
invertebrates, and fish. Acid-neutralizing capacity (ANC) is a measure of the buffering
capacity of natural waters against acidification. Even though ANC does not directly alter
the health of biota, it is a key metric of acidification that relates to pH and aluminum.
Biological effects are primarily attributable to low pH and high inorganic aluminum
concentration. Characterization of ANC and its levels of concern have not changed
appreciably with the newly available information since the 2008 ISA. Few or no fish
species are found in lakes and streams that have very low ANC (near zero) and low pH
(near 5.0), and the number of fish species generally increases with higher ANC and pH
(Appendix 8.3). The fish lost to acidification include culturally and recreationally
important species.
Acidified terrestrial habitats are characterized by detrimental physiological effects to
vegetation, including inhibited growth and decreased plant health. Acidifying deposition
can decrease membrane stability and freezing tolerance in young red spruce needles. For
many species, calcium depletion from the soil and aluminum mobilization cause
decreased root uptake of calcium and disrupt fine root physiological functions. Reduced
availability of cations in the soil can also make trees more vulnerable to other stresses,
such as damage from insects and other pathogens. Within the eastern U.S., the
physiological effects of acidifying deposition have been well documented for several
culturally important tree species with known ecosystem services, particularly sugar maple
(Acer saccharum) and red spruce (Picea rubens). Evidence available before and since the
2008 ISA shows that there is consistent and coherent evidence among these species that
acidifying deposition can decrease foliar cold tolerance, increase rates of crown dieback,
decrease tree growth, suppress seedling regeneration, and increase mortality rates
(Chapter 1, Section 1.5.3; Appendix 5). Since the 2008 ISA, studies from the northeastern
U.S. have shown that calcium addition can alleviate many of these effects, demonstrating
that acidification effects can be ameliorated in the short-term by soil amendments and
adding to the evidence that there is potential for recovery. Acidifying deposition has also
been linked to changes in forest understory plant community composition in the
northeastern U.S., grass and forb biodiversity in eight ecoregions across the U.S., and
decreased grassland plant species richness in Europe.
lxxv
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Nitrogen (N) Enrichment/Eutrophication of Terrestrial, Wetland, and Aquatic
Ecosystems
Terrestrial, wetland, freshwater, and estuarine ecosystems in the U.S. are affected by N
enrichment/eutrophication caused by N deposition. N enrichment/eutrophication refers to
N nutrient-driven changes in growth, physiology, and biodiversity. These effects have
been consistently documented across the U.S. for hundreds of species. New evidence
strengthens the causal relationships for ecosystem N enrichment/eutrophication
determined in the 2008 ISA (Chapter 1, Table 1-1). and several new causal relationships
have been identified.
The 2008 ISA documented that in sensitive terrestrial and wetland ecosystems, the N
enrichment effect starts with the accumulation of N in the soil. This increases the
availability of N, a nutrient that increases the growth of some species of soil microbes
and vascular plants at the expense of other species, which may decrease biodiversity.
Since the 2008 ISA, the largest increase in ecological evidence is for terrestrial N driven
enrichment/eutrophication effects (Chapter 1, Section 1.5.1. Section 1.5.2; Appendix 4.
and Appendix 6). This new research confirms the causal relationship between N
deposition and ecological effects documented in the 2008 ISA and improves our
understanding of the mechanistic links that inform causal determinations between N
deposition, biogeochemistry, and biota in terrestrial ecosystems (Chapter 1, Table 1 -1). A
new causal determination has been added to reflect more specific categories of effects to
include physiology, growth, and ecosystem productivity. Further, there is now stronger
empirical evidence from across most regions of the U.S. to quantify the levels of N
deposition (empirical critical loads [CLs]) that cause a myriad of ecological effects,
including: shifts in community composition and biodiversity declines in lichens and
grasses/forbs; tree growth and mortality; and increased nitrate leaching (Chapter 1,
Figure 1-7).
Since the 2008 ISA, studies have strengthened the findings of N effects on decreasing
lichen and mycorrhizal fungi biodiversity and provided additional CL estimates. In
terrestrial ecosystems, new evidence supports that epiphytic lichens (an algal and or
cyanobacteria-fiingal symbiont) and mycorrhizae (a plant-fungal symbiosis at the tips of
plant roots) are the organisms that are most sensitive to atmospheric N deposition and
acidifying deposition. Although lichens typically are only a small portion of terrestrial
biomass, these changes in lichen communities are meaningful because lichens provide
food and habitat for insects, birds, and mammals; contribute to nutrient and hydrologic
cycling; have many traditional human uses; and have considerable potential for
pharmaceutical use. Changes in the community composition of mycorrhizal fungi and
declines in mycorrhizal abundance have been observed in the U.S. These fungi are
lxxvi
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
important for supplying nutrients and water to plants, influencing soil C sequestration,
and producing fruiting bodies (mushrooms) used by humans and wildlife.
The effects of N deposition on tree growth are well documented. There is now evidence
of widespread species-specific effects of N deposition on tree growth and mortality in the
U.S. While tree growth has generally been enhanced by N deposition over the last several
decades, there is wide variation among species in mortality and growth in response to N
deposition. This information, however, has not been used yet to quantitatively assess
changes in tree community composition.
In wetlands, the 2008 ISA documented that wetlands receiving a larger fraction of their
total water budget in the form of precipitation are more sensitive to the effects of N
deposition. For example, bogs and fens (55-100% of hydrological input from rainfall) are
more sensitive to N deposition than coastal wetlands (10-20% as rainfall). Since the 2008
ISA, CLs for U.S. coastal and freshwater wetlands have been established. The CL for
freshwater wetlands is based on C cycling, as well as biodiversity represented by the
morphology and population dynamics of the purple pitcher plant (Sarracenia purpurea).
The CL for coastal wetlands is based on several different ecological endpoints, including
plant community composition, microbial activity, and biogeochemistry.
The 2008 ISA documented that the process of N eutrophication is similar in freshwater
and estuarine ecosystems and typically begins with a nutrient-stimulated algal bloom that
is followed by anoxic conditions. The lack of oxygen in the water due to the respiration
and decomposition of the algae affects higher tropic species. The contribution of N
deposition to total N loading varies among freshwater lakes and stream ecosystems.
Atmospheric deposition is the main source of new N to most headwater streams,
high-elevation lakes, and low-order streams far from the influence of other N sources like
agricultural runoff and wastewater effluent. N deposition was known at the time of the
2008 ISA to alter biogeochemical processes and nutrient ratios in recipient freshwater
ecosystems. New CLs support previous observations of increased productivity of
phytoplankton and algae, species changes, and reductions in diversity in atmospherically
N enriched lakes and streams. The productivity of many freshwater ecosystems is N
limited. Thus, even small amounts of N can shift nutrient ratios and affect the trophic
status of lakes and streams. As reported in the 2008 ISA and newer studies, a shift from
N limitation to either colimitation by N and P or limitation by P has been observed in
some alpine lakes in the U.S. and other countries, with these shifts correlated with
elevated N deposition.
At the time of the 2008 ISA, N was recognized as the major cause of harm to the majority
of estuaries in the U.S. Since 2008, new paleontological studies, observational studies,
and experiments have further characterized the effects of N on phytoplankton growth and
lxxvii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
community dynamics, macroinvertebrate response, and other indices of biodiversity. For
this ISA, new information is consistent with the 2008 ISA, and the causal determination
has been updated to reflect more specific categories of effects to include total primary
production, altered growth, and total algal community biomass.
Since the 2008 ISA, studies have shown that coastal acidification can be exacerbated by
elevated N input. With increasing N inputs to coastal waters, CO2 in the water column is
produced from the degradation of excess organic matter, as well as respiration of living
algae and seagrasses, which in turn can make the water more acidic. Coastal acidification
is projected to alter marine habitat, have a wide range of effects at the population and
community level, and impact food web processes. Newer studies show that organisms
that produce calcium carbonate shells are impacted by increasing acidification of ocean
waters. Decreased concentration of carbonate ions (which organisms such as calcareous
plankton, oysters, clams, sea urchins, and corals absorb to build shells) are observed in
acidic conditions. Changes in the carbonate system, including decreased pH, have been
shown to elicit biological responses in commercially important species from the New
England coast, and documented declines of oyster production on the U.S. West Coast are
linked to ocean acidification. Research on stressors associated with conditions of coastal
acidification and eutrophication suggests that interactions between elevated CO2,
decreasing pH, and nutrient inputs are complex.
Sulfur (S) Enrichment on Wetland and Freshwater Ecosystems
SOx deposition increases SO42 concentration in surface waters. New evidence supports
links between aqueous sulfur concentrations in freshwater ecosystems and both mercury
methylation and sulfide toxicity (Table ES-1): however, quantitatively linking these
outcomes to atmospheric deposition remains a challenge.
Increasing SO42 concentration in surface waters can stimulate the microbial
transformation of inorganic Hg into methylmercury (MeHg; Appendix 12). MeHg is the
most persistent and toxic form of Hg affecting animals in the natural environment.
Indicators of S deposition effects upon Hg methylation include increases in MeHg
concentrations or fraction of total Hg in water, sediments, and peat, as well as increases
in MeHg concentrations in periphyton, submerged aquatic plants, invertebrates, and fish.
New evidence confirms the relationship between aqueous concentrations of SO42 and
MeHg, and broadens our understanding of where methylation occurs from the wetlands
and lakes reported in the 2008 ISA to include rivers, reservoirs, streams, and saturated
forest soils. Hg methylation occurs at anoxic-oxic boundaries in peat moss and
periphyton, as well as in wetland, lake, estuarine, and marine sediments. There are
published quantitative relationships between surface water SO42 concentrations and
MeHg concentrations, MeHg and total Hg in water, and Hg load in larval mosquitoes and
lxxviii
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
fish. There is also evidence that decreasing sulfur deposition loads over time
(observational studies of SOx deposition, experimental studies of simulated SOx wet
deposition) result in lower concentrations of MeHg in water, invertebrates, and fish.
There is new evidence since the 2008 ISA to infer a causal relationship between S
deposition and sulfide phytotoxicity, which alters growth and productivity, species
physiology, species richness, community composition, and biodiversity in wetland and
freshwater ecosystems (Appendix 12). This new causal statement reflects new research
on sulfide phytotoxicity in North American wetlands, as the 2008 ISA described sulfide
phytotoxicity only in European ecosystems. Current levels of S deposition cause sulfide
toxicity in wetland and aquatic plants. Indicators of sulfide phytotoxicity caused by S
deposition include increases in water or sediment sulfide concentrations. Sulfide
negatively effects growth, competition, and persistence in several wetland species,
including the economically important species wild rice, and the keystone sawgrass
species in the Everglades marshes. To date, no published studies have established
regional sensitivities to sulfide phytotoxicity, although studies have observed its effects in
New York, Minnesota, and Florida freshwater marshes. There are no S deposition-based
critical loads for mercury methylation or sulfide phytotoxicity, although researchers have
proposed water quality values to protect biota against these effects in several ecosystems
(Appendix 12).
Ecological Effects of Particulate Matter (PM) Other Than Those Associated with
Nitrogen (N) and Sulfur (S) Deposition
There is a likely causal relationship between PM and ecological effects on biota, other
than those associated with N and S deposition (Table ES-1; Appendix 15). Since
publication of the 2009 PM ISA, new literature has built upon the existing knowledge of
ecological effects associated with PM components, especially metals and organics. In
some instances, new techniques have enabled further characterization of the mechanisms
of PM on soil processes, vegetation, and effects on fauna. New studies provide additional
evidence for community-level responses to PM deposition, especially in soil microbial
communities. However, uncertainties remain due to the difficulty in quantifying
relationships between ambient concentrations of PM and ecosystem response.
Ecosystem Services
The ecosystem services literature has expanded since the 2008 ISA to include studies that
better characterize ecosystem service valuation and quantification related to acidification
and N enrichment/eutrophication.
New valuation studies for ecosystem acidification pair biogeochemical modeling and
benefit transfer equations informed by willingness-to-pay surveys, especially for the
June 2018
lxxix
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Adirondacks and Shenandoah regions (Appendix 14). Despite this progress, for many
regions and specific services, poorly quantified relationships between deposition,
ecological effect, and services are the greatest challenge in developing specific data on
the economic benefits of emission reductions.
In the 2008 ISA, there were no publications that had specifically evaluated the effects of
N deposition on ecosystem services associated with N driven enrichment/eutrophication.
Since the 2008 ISA, several comprehensive studies have been published on the
ecosystems services related to N pollution in the U.S. These include an evaluation of
services affected by multiple N inputs (including N deposition) to the Chesapeake Bay, a
synthesis of the cost-benefits on N loading across the nation, and an analysis of how N
leaked from its intended area of application (e.g., agricultural fields) affects ecosystem
services of adjacent ecosystems. Most notably, new work identifies 1,104 unique chains
linking N deposition to human beneficiaries.
Considering the full body of literature on ecosystem services related to N and S, the
following conclusions are offered: (1) there is evidence that N and S emissions/deposition
have a range of effects on U.S. ecosystem services and their social value; (2) some
economic studies demonstrate such effects in broad terms, but it remains
methodologically difficult to derive economic costs and benefits associated with specific
regulatory decisions/standards; and (3) over a thousand relationships are now
documented between N and S air pollution and changes in final ecosystem goods and
services.
Integrating across Ecosystems
Overall, new evidence since the 2008 ISA increases the weight of evidence for ecological
effects, confirming concepts previously identified and improving quantification of
dose-response (or deposition-ecological indicator) relationships, particularly for N and S
deposition. The ecological effects are described by the causality determinations
(Figure ES-3). Figure ES-3 reorganizes the information in Figure ES-2 to offer a
visualization of the effects of NOy, SOx, and PM by ecosystem type (e.g., terrestrial,
wetland, freshwater, and estuarine). The gas-phase effects were not included in this
diagram. With this organization, the multiple effects occurring in each ecosystem due to
various pollution combinations of NOy, SOx, and PM are emphasized. Between two and
four different classes of pollutant effect may occur in each ecosystem type in the U.S. For
more information on key messages see the expanded discussion in Chapter 1; detailed
information on specific ecosystem types and specific classes of pollutant effects included
in the ISA may be found in the Appendices.
June 2018
lxxx
DRAFT: Do Not Cite or Quote

-------
Causal
Likely causal
Suggestive
Inadequate
Not likely
Not evaluated In causal framework
Wetland
Fresh water
Estuaiy
dau of Pollutant Effect
W-emichfneiYt/
EutropWcatton
N-ewlchment/
EutropWcatlon
Sulfide
Toxicity
Mercury
Melhylatioo
Aerification
W-ewlchment/
Evtrop*iJcatl©n
Sutflde
Toxicity
Mercury
Methylatlon
N-ewlcSwnent/
Etftrojrfwcatkwt
Eutropfckatton
driven
acidification
Causality framework
Ecosystem
Terrestrial
Direct
phrytotoxk
Acidification
Indicator
Productivity
Biodiversity
Growth rate
Sal or sediment
chemistry
Surface water
chemistry
Gases I
N*S dep
N+Sdep
N dep
Physiological
ateration, stress or
* Includes: NO, N02. HNOj, SOa, and PAN
Figure ES-3 Causal relationships between the criteria pollutants and ecological effects organized under
ecosystem type.
June 2018
lxxxi
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
CHAPTER 1
INTEGRATED SYNTHESIS
1.1 Introduction to This Integrated Science Assessment (ISA)
1.1.1 Purpose
The Integrated Science Assessment (ISA) is a comprehensive evaluation and synthesis of
the policy-relevant science. Policy-relevant science is that which is "useful in indicating
the kind and extent of all identifiable effects on public health or welfare which may be
expected from the presence of [a] pollutant in the ambient air," as described in
Section 108 of the Clean Air Act (C'AA. 1990a).1 This ISA communicates critical science
judgments of the ecological criteria for oxides of nitrogen, oxides of sulfur, and
particulate matter. Accordingly, this ISA is the scientific foundation for the review of the
ecological effects of the current secondary (welfare-based) National Ambient Air Quality
Standards (NAAQS) for oxides of nitrogen, oxides of sulfur, and particulate matter. The
Clean Air Act definition of welfare effects includes, but is not limited to, effects on soils,
water, wildlife, vegetation, visibility, weather, and climate, as well as effects on
man-made materials, economic values, and personal comfort and well-being. The
nonecological welfare effects associated with particulate matter, such as climate and
visibility, are considered part of a separate, ongoing review of PM that is outlined in the
Integrated Review Plan for the National Ambient Air Quality Standards for Particulate
Matter (IRP) (U.S. EPA. 2016d).Thc human health effects for oxides of nitrogen, oxides
of sulfur, and particulate matter are evaluated in separate assessments conducted as part
of the review of the primary (human health-based) NAAQS for oxides of nitrogen (U.S.
EPA. 2016f). oxides of sulfur (U.S. EPA. 2016e). and as noted above, in the ongoing
review for particulate matter (U.S. EPA. 2016dV
Oxides of nitrogen, oxides of sulfur, and particulate matter (PM) are reviewed here
together because they are inter-related through complex chemical and physical
atmospheric processes and because they all contribute to nitrogen (N) and sulfur (S)
deposition, which in turn contributes to well-documented ecological effects. In this
document, the term "oxides of nitrogen" refers to all forms of oxidized nitrogen (N)
compounds, including NO, NO2, and all other oxidized N containing compounds formed
1 The general process for developing an ISA, including the framework for evaluating weight of evidence and
drawing scientific conclusions and causal judgments, is described in a companion document, Preamble to the
Integrated Science Assessments (U.S. EPA. 2015e').
June 2018
1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
from NO and NO2.1 Oxides of sulfur2 are defined here to include sulfur monoxide (SO),
sulfur dioxide (SO2), sulfur trioxide (SO3), disulfur monoxide (S2O), and sulfate (SO42 ).
However, SO, SO3, and S2O are present at much lower ambient levels than SO2 and SO42
and are therefore not discussed further. Particulate matter is composed of some or all of
the following components: nitrate (NO3 ). SO42 . ammonium (NH4+), metals, minerals
(dust), and organic and elemental carbon (C).
This ISA updates the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur—Ecological Criteria [hereafter referred to as the 2008 ISA (U.S. EPA. 2008a)l. as
well as the ecological portion of the Integrated Science Assessment for Particulate Matter
(U.S. EPA. 2009a). with studies and reports published from January 2008 through May
2017. Thus, this ISA updates the state of the science that was available for the 2008 ISA,
which informed decisions on the secondary oxides of nitrogen and oxides of sulfur
NAAQS in the review completed on March 20, 2012. In the final rulemaking, the
Administrator's decision was that, while the current secondary standards were inadequate
to protect against adverse effects from deposition of oxides of nitrogen and oxides of
sulfur, it was not appropriate under Section 109(b) to set any new secondary standards at
this time due to the limitations in the available data and uncertainty as to the amount of
protection the metric (Aquatic Acidification Index—see Chapter 1.2.2.6) developed in
the Policy Assessment (U.S. EPA. 2011a) would provide against acidification effects
across the country (77 FR 20281). In addition, the Administrator decided that it was
appropriate to retain the current nitrogen dioxide (NO2) and sulfur dioxide (SO2)
secondary standards to address direct effects of gaseous NO2 and SO2 on vegetation.
Thus, taken together, the Administrator decided to retain and not revise the current NO2
and SO2 secondary standards: an NO2 standard set at a level of 0.053 ppm, as an annual
arithmetic average, and an SO2 standard set at a level of 0.5 ppm, as a 3-hour average, not
to be exceeded more than once per year (77 FR 20281). The current secondary standards
for PM are intended to address PM-related related welfare effects, including visibility
impairment, ecological effects, effects on materials, and climate impacts. These standards
are a 3-year annual mean PM2 5 concentration of 15 |ig/m3. with the 24-hour average
PM2 5 and PM10 set at concentrations of 35 |ig/m3 and 150 |ig/m3. respectively.
1	This ISA reserves the abbreviation NOx strictly as the sum of NO and NO2—consistent with that used in the
atmospheric science community—and uses the term "oxides of nitrogen" to refer to the broader list of oxidized
nitrogen species. Oxides of nitrogen refers to NOy as the total oxidized nitrogen in both gaseous and particulate
forms. The major gaseous and particulate constituents of NOy include nitric oxide (NO), nitrogen dioxide (NO2),
nitric acid (HNO3), peroxyacetyl nitrate (PAN), nitrous acid (HONO), organic nitrates, and particulate nitrate (NO3).
This ISA uses the definitions adopted by the atmospheric sciences community.
2	Oxides of sulfur refers to the criteria pollutant category.
June 2018
2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
This new review of the secondary oxides of nitrogen, oxides of sulfur, and particulate
matter NAAQS is guided by several policy-relevant questions that were identified in The
Integrated Review Plan for the Secondary National Ambient Air Quality Standard for
Nitrogen Oxides, Sulfur Oxides, and Particulate Matter [hereafter referred to as the 2017
IRP, (U.S. EPA. 2017b)!.
To address these questions, this ISA aims to characterize the evidence available in the
peer-reviewed literature for ecological effects associated with:
•	the major gaseous and particulate constituents of total oxidized N (NOy), which
include NO, NO2, HNO3, PAN, HONO, organic nitrates, and NO3 :
•	the major gaseous and particulate constituents of SOx, which include SO2 and
SO42 ; and
•	PM.
The assessment activities include:
•	Identifying policy-relevant literature.
•	Evaluating strength, limitations, and consistency of findings.
•	Integrating findings across scientific disciplines and across related ecological
outcomes.
•	Considering important uncertainties identified in the interpretation of the scientific
evidence.
•	Assessing policy-relevant issues related to quantifying ecological risks, such as
ambient air concentrations, deposition, durations, and patterns associated with
ecological effects; the relationship between ambient air concentrations, deposition,
and ecological response and the existence of thresholds below which effects do
not occur; and species and populations potentially at increased risk of ecological
effects.
New analyses with the goal of quantifying risk, such as new model runs, critical loads
exceedance maps, and quantified uncertainties regarding modeled scenarios are not
conducted in the ISA. These are types of analyses, if pursued, require the selection of
chemical or biological limits that define critical loads and represent adversity. These
analyses would also require choosing a time period over which to average deposition.
Such scope-of-analysis decisions are more appropriate for the Risk and Exposure
Assessment, as described in the 2017 IRP (U.S. EPA. 2017b). The information
summarized in this ISA will serve as the scientific foundation of the Risk and Exposure
and Policy Assessments during the current review of the secondary oxides of nitrogen,
oxides of sulfur, and particulate matter NAAQS.
3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
1.1.2
Process and Development
The U.S. EPA uses a structured and transparent process to evaluate scientific information
and determine the causality of relationships between air pollution and ecological effects
[see Preamble (U.S. EPA. 20156)1. The ISA development includes approaches for
literature searches, criteria for selecting and evaluating relevant studies, and a framework
for evaluating the weight of evidence and forming causal determinations. As part of this
process, the ISA is reviewed by the public and by the Clean Air Scientific Advisory
Committee (CASAC), which is a formal independent panel of scientific experts. This ISA
informs the review of the secondary oxides of nitrogen, oxides of sulfur, and particulate
matter NAAQS and therefore integrates and synthesizes information characterizing NOy,
SOx, and PM air concentrations. It also examines deposition of these substances and their
ecological effects. Relevant studies include those examining atmospheric chemistry,
spatial and temporal trends, and deposition, as well as U.S. EPA analyses of air quality
and emissions data. Relevant ecological research includes geochemistry, microbiology,
physiology, toxicology, population biology, and community ecology. The research
includes laboratory and field additions, as well as gradient studies.
The U.S. EPA conducted literature searches to identify relevant peer-reviewed studies
published since the previous ISA (i.e., from January 2008 through December 2015;
Chapter 1, Figure 1-1).
June 2018
4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
HERO = Health and Environmental Research Online; ISA = Integrated Science Assessment.
Figure 1-1 Workflow for collecting relevant literature for the 2017 Integrated
Science Assessment for Oxides of Nitrogen, Oxides of Sulfur, and
Particulate Matter—Ecological Criteria.
Multiple search methods were used in the Web of Science database [Preamble (U.S.
EPA. 20156). Appendix 21. including searches by keyword and by citations of 2008 ISA
references. Subject-area experts and the public were also permitted to recommend studies
and reports during kick-off workshops held at the U.S. EPA in March 2014 for oxides of
nitrogen and oxides of sulfur and in February 2015 for particulate matter. The new
references were sorted by automated methods into topic areas based on wording in the
publication's abstract or numbers of citations of 2008 ISA references, and the resultant
relevant literature was reviewed by the ISA authors. Studies were screened based on the
title first and then by the abstract; studies that did not address a relevant research topic
based on this screening were excluded. The U.S. EPA also identified studies from
previous assessments as definitive works on particular topics to include in this ISA. The
HERO project page for this ISA (http://hero.epa.aov/heronet/NOYSOYPMEco') contains
the references that are cited in the ISA and electronic links to bibliographic information
and abstracts.
June 2018
5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
The Preamble to the Integrated Science Assessments (U.S. EPA. 2015e) describes the
general framework for evaluating scientific information, including criteria for assessing
study quality and developing scientific conclusions. For ecological studies, emphasis is
placed on studies that characterize quantitative relationships between criteria pollutants
and ecological effects that occur at concentration and deposition levels relevant to current
ambient levels in the U.S. However, experimental studies with higher exposure
concentrations are included if they contribute to an understanding of mechanisms.
This ISA draws conclusions about relationships between NOy, SOx, and PM and
ecological effects by integrating information across scientific disciplines and related
ecological outcomes and synthesizing evidence from previous and recent studies.
Determinations are made about causation, not just association, and are based on
judgments of consistency, coherence, and scientific plausibility of observed effects, as
well as related uncertainties. The ISA uses a formal causal framework [Table II of the
Preamble (U.S. EPA. 2015eYI which is based largely on the aspects for causality
proposed by Sir Bradford Hill to classify the weight of evidence according to the
five-level hierarchy summarized below.
•	Causal relationship
•	Likely to be a causal relationship
•	Suggestive of, but not sufficient to infer, a causal relationship
•	Inadequate to infer the presence or absence of a causal relationship
•	Not likely to be a causal relationship
1.1.3 Organization
This ISA includes the Preface (legislative requirements and history of the secondary
oxides of nitrogen, oxides of sulfur, and particulate matter NAAQS), an Executive
Summary, an Integrated Synthesis chapter, and 16 appendices. The general process for
developing an ISA is described in a companion document, Preamble to the Integrated
Science Assessments (U.S. EPA. 2015e). Chapter 1 synthesizes the scientific evidence
that best informs policy-relevant questions that frame this review. Appendix 1 is an
introduction to the appendices. Appendix 2 characterizes the sources, atmospheric
processes, and the trends in ambient concentrations and deposition of NOy, SOx, and PM.
Appendix 3 describes direct effects of NOy and SOx gases on plants and lichens.
Appendix 4-Appendix 6 describe N and S deposition effects on terrestrial
biogeochemistry and the terrestrial biological effects of terrestrial acidification and N
enrichment. Appendix 7 describes the effects of N and S deposition on aquatic
biogeochemistry. Appendix 8-Appendix 10 characterize the biological effects of
June 2018
6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
freshwater acidification, freshwater N enrichment, and N enrichment in estuaries and
near-coastal systems. Appendix 11 describes the effects of N deposition on wetlands, and
Appendix 12 characterizes the ecological effects of S as a nutrient. Appendix 13 presents
information on climate modification of ecosystem response to N and S, while ecosystem
services are discussed in Appendix 14. Appendix 15 is a review of the ecological effects
of forms of PM, which are not related to N or S deposition. Finally, Appendix 16 presents
case studies for six locations in the U.S. (southern California, northeastern U.S., Rocky
Mountain National Park, southeastern Appalachia, Tampa Bay, and the Adirondacks)
where data are sufficient to well characterize the ecological effects of N and S deposition.
These sites would therefore make good candidates for further understanding the linkages
across various effects and ecosystems in a location to better assess risk and exposure.
1.2 Connections, Concepts, and Changes
1.2.1 Connections
Although scientific material in this ISA is divided into separate appendices for
atmospheric science and the multiple ecological effects, the strong links between the
atmosphere and terrestrial and aquatic ecosystems are acknowledged (Chapter 1,
Figure 1-2). Emissions of NOy, SOx, and PM cause an accumulation of N and S in the
environment that creates a multitude of effects on terrestrial, wetland, and aquatic
ecosystems. Nitrogen is a vital component of all biological systems, serving as an
essential element to molecules such as amino acids and nucleic acids, which are among
the biochemical building blocks of life. As an organizing concept to understand the
effects of N within the environment, the sequence of transfers, transformations, and
environmental effects has been described as the "N cascade" (Galloway and Cowling.
2002). The concept of cascading effects also applies to S, which is also an essential
macronutrient.
June 2018
7
DRAFT: Do Not Cite or Quote

-------
Ecosystem Services
Appendix 14
Biological Effects
of S-riutrient
Appendix 12
Gas-phase
Ecological Effects
Appendix 3
Oxidation
so2—»-h2so4
NOx—>¦ HNOj
Dissolution
2H++ S042"
->-H++N0r
Other Ecological
Effects of PM
Appendix 15
Dry deposition
NO ,NH , SO
Biological Effects of
Terrestrial Acidification
Appendix 5
Biological Effects of
Terrestrial N Enrichment
Appendix 6
n2o no,
Wetlands
Appendix 11
Soil BGC
Appendix 4
Aquatic BGC
Appendix 7
Wet deposition
H+, NH4+, N03", SQ,2"
Biological Effects of
Freshwater Acidification
Appendix 8
Biological Effects
of Freshwater j
N Enrichment
Appendix 9
Biological Effects of
Estuarine N Enrichment
Appendix 10
Deposition
Ecological
Effect
Atmospheric Sciences
Appendix 2
Climate Modification
of Ecological Effects
Appendix 13
Ambient Air
Concentration
Ca2+ = calcium ion; GHG = greenhouse gas; H* = hydrogen ion; HN03 = nitric acid; H2S04 = sulfuric acid; Mg2+ = magnesium ion;
N = nitrogen; N20 = nitrous oxide; NH3 = ammonia; NH/ = ammonium; NHX = NH3 + NH4+ + reduced organic nitrogen compounds;
NO = nitric oxide; N02 = nitrogen dioxide; N03" = nitrate; NOx = NO + N02; PAN = peroxyacetyl nitrate; PM = particulate matter;
S02 = sulfur dioxide; S042" = sulfate; SOx = S02 + S042"; VOC = volatile organic compounds.
The sum of reactive oxidized nitrogen species is referred to as NOY (NOY = NO + N02 + HN03 + 2N205 + HONO + N03" + N20
PAN + other organic nitrates). VOC refers to volatile organic compounds.
Although not explicitly indicated, wet and dry deposition of PM components (e.g., metals, minerals, and secondary organic aerosol)
also occur and contribute to ecological effects.
Source: Modified from U.S. EPA (20083).
Figure 1-2 Overview of atmospheric chemistry, deposition, and ecological
effects of emissions of oxides of nitrogen, oxides of sulfur, and
reduced nitrogen.
1.2.2 Concepts
1	This ISA draws on many methodological approaches and disciplines within the larger
2	scientific fields of ecology and atmospheric sciences. The studies discussed herein are
3	best understood in the context of some general concepts within these fields, such as
4	ecosystem scale, structure, and function (Chapter 1.2.2.1); deposition and source
5	apportionment to ecosystems (Chapter 1.2.2.2k critical loads [CL, (Chapter 1.2.2.3)1;
6	biodiversity (Chapter 1.2.2.411; the effects of reduced versus oxidized forms of N
June 2018
8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
(Chapter 1.2.2.5); and the metric developed in the previous secondary NAAQS review,
the Aquatic Acidification Index [AAI, (Chapter 1.2.2.6)1.
Ecosystem structure comprises both biodiversity and geography. Biodiversity
encompasses many quantitative measures of the abundance and distribution of organisms
within a defined geographical area (for a more explicit definition, see Chapter 1.2.2.1 and
Chapter 1.2.2.4). Ecosystem function refers to processes that control fluxes and pools of
matter and energy in the ecosystem (Chapter 1.2.2. IV The loss of biodiversity is a key
consequence of the air pollutants discussed in this ISA. The importance of preserving
biodiversity and ecosystem function contributes to the sustainability of ecosystem
services that benefit human welfare and society (Chapter 1.2.2.4 and Appendix 14).
In human health assessments, dose-response relationships are used to identify
quantitative relationships between chemical exposure (dose) and health outcomes
(response), with emphasis on identifying thresholds, the lowest doses at which negative
health outcomes are observed. In ecology, CLs provide a similar quantitative relationship
between chemical dose (e.g., deposition) and specific, quantitative changes in ecological
properties or processes (Chapter 1.2.2.3). Use of CLs in evaluating the effects of
deposition upon ecosystems must consider how deposition compares to other
anthropogenic and ambient sources of N and S to these ecosystems (Chapter 1.2.2.2). as
well as the heterogeneous sensitivities of organisms and ecosystems to different chemical
forms of deposition (Chapter 1.2.2.5).
1.2.2.1 Ecosystem Scale, Structure, and Function
For this assessment, an ecosystem is defined as the interactive system formed from all
living organisms (biota) and their abiotic (chemical and physical) environment within a
given area (IPCC. 2007a). Ecosystem spatial boundaries are somewhat arbitrary,
depending on the focus of interest or study. Thus, the spatial extent of an ecosystem may
range from very small, well-circumscribed systems such as a small pond, to biomes at the
continental scale, or the entire globe (U.S. EPA. 2008a). Ecosystem spatial scale does not
always correlate with complexity. A small pond may be a complex system with multiple
trophic levels ranging from phytoplankton to invertebrates to several feeding guilds of
fish. A large lake, on the other hand, may be a very simple ecosystem, such as the Great
Salt Lake in Utah that covers approximately 1,700 square miles but contains only
bacteria, algae, diatoms, and two invertebrate species (U.S. EPA. 2013b). All ecosystems,
regardless of size or complexity, have multiple interactions between biota and abiotic
factors and a reduction in entropy through energy flow from autotrophs to top predators.
Ecosystems include both structural (geography and biodiversity; e.g., soil type and food
June 2018
9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
web trophic levels) and functional (flow of energy and matter; e.g., decomposition,
nitrification) attributes. Ecosystem changes are often considered undesirable if important
structural or functional components of the ecosystems are altered following pollutant
exposure (U.S. EPA. 2013b. 1998a).
Biotic or abiotic structure may define an ecosystem. Abiotic structure includes climatic
and edaphic components. Biological structure includes species abundance, richness,
distribution, evenness, and composition, measured at the population, species, community,
ecosystem, or global scale. A species (for eukaryotic organisms) is defined by a common
morphology, genetic history, geographic range of origin, and ability to interbreed and
produce fertile offspring. A population consists of interbreeding groups of individuals of
the same species that occupy a defined geographic space. Interacting populations of
different species occupying a common spatial area form a community (Barnthousc ct al..
2008). Community composition may also define an ecosystem type, such as a pine forest
or a tall grass prairie. Pollutants can affect the ecosystem structure at any of these levels
of biological organization (Sutcr et al.. 2005).
Individual plants or animals may exhibit changes in metabolism, enzyme activities,
hormone function, or may suffer gross lesions, tumors, deformities, or other pathologies.
However, only some organism-level endpoints such as growth, survival, and reproductive
output have been definitively linked to effects at the population level and above (U.S.
EPA. 2013b). Population-level effects of pollutants include changes overtime in
abundance or density (number of individuals in a defined area), age or sex structure, and
production or sustainable rates of harvest (Barnthouse et al.. 2008). Community-level
attributes affected by pollutants include species richness, species abundance,
composition, evenness, dominance of one species over another, or size (area) of the
community (U.S. EPA. 2013b). Pollutants may affect communities in ways that are not
observable in organisms or populations (Battel 1. 2007). including (1) effects resulting
from interactions between species, such as altering predation rates or competitive
advantage; (2) indirect effects, such as reducing or removing one species from the
assemblage and allowing another to emerge (Petraitis and Latham. 1999); and
(3) alterations in trophic structure.
Alternatively, ecosystems may be defined on a functional basis. "Function" refers to the
suite of processes and interactions among the ecosystem components that involve energy
or matter. Examples include water dynamics and the flux of trace gases such as rates of
photosynthesis, decomposition, nitrification, or carbon cycling. Pollutants may affect
biotic structure indirectly. For example, a pollutant may first alter abiotic conditions
(e.g., soil chemistry), which in turn influences biotic structure and function (Bartell.
2007).
10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Some ecosystems, and some aspects of particular ecosystems, are less vulnerable to
long-term consequences of pollutant exposure. Other ecosystems may be profoundly
altered if a single attribute is affected. Thus, spatial and temporal definitions of
ecosystem structure and function become an essential factor in defining impacted
ecosystem services and CLs of particular pollutants, either as single pollutants or in
combination with other stressors.
The main causal determinations of this ISA (Chapter 1.2.3) are that N and S deposition
affect ecosystem structure, with effects ranging from biogeochemical alterations in soil
and water chemistry to multiple levels of biological organization, including species-level
alterations of physiological processes, shifts in biodiversity and ecological function.
1.2.2.2 Deposition and Source Apportionment of Nitrogen (N) and Sulfur (S) to
Ecosystems
Deposition of N and S results from a variety of human activities and atmospheric
processes. Emissions from stationary, mobile, and agricultural sources undergo
atmospheric transformation (Chapter 1.3.1) to form products that are more readily
deposited out of the air onto the land or waterscape (Chapter 1.3.3). The contribution of
atmospheric deposition to total loading for N and S varies within and among terrestrial,
wetland, freshwater, and estuarine ecosystems.
In the 2008 ISA, atmospheric deposition was identified as the main source of
anthropogenic N to unmanaged terrestrial ecosystems. This conclusion has been
confirmed by new studies on N sources to lands and waterways, which find
human-mediated watershed N inputs that range from <1.0 to 34.6 times the rate of
background N input (Appendix 4.2). Across all watersheds, atmospheric N deposition is
the second largest overall human-mediated N source and the largest N source to 33% of
watersheds. Current deposition levels in the U.S. are discussed in Appendix 2 and
Chapter 1.3.3. No new information has been published on nonatmospheric sources of S in
terrestrial ecosystems (Appendix 4.2); S inputs from the atmosphere are discussed in
Appendix 2 and Chapter 1.3.3.
In the 2008 ISA, atmospheric deposition was also identified as the main source of N to
some freshwater ecosystems, including headwater streams, high-elevation lakes, lower
order streams in undisturbed areas and freshwater wetlands (e.g., bogs and fens).
Evidence for the influence ofN deposition on water chemistry has been further supported
by new studies that quantify the contribution of N deposition to total N loading in
freshwater lakes and streams, and which quantify atmospheric contributions during storm
events (Table 7-1). As shown in these studies, deposition can represent a substantial
June 2018
11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
portion of total N loading to surface waters. However, other nonpoint and point sources
of N dominate N inputs to high-order streams.
In fresh surface waters and wetlands, sources of S that contribute to enrichment effects
are the same sources of S that induce acidifying effects; these sources include weathering
of minerals in sediments and rocks, leaching from terrestrial S cycling, internal cycling,
and direct atmospheric deposition. The 2008 ISA showed that drought can release S
stored in wetlands or lake sediments, as bound sulfide (S2 ) is exposed to atmospheric
oxygen and oxidized to sulfate. Model of Acidification of Groundwater in Catchments
(MAGIC) modeling has shown that variable water levels due to climate change-induced
droughts can increase water sulfate concentrations, and observational research has shown
that drought increases lake S load by 5 kg S/ha/yr. New evidence confirms that
fluctuating water levels in wetlands increase SO42 concentration in pulses following
water level recovery.
Sources of N in coastal areas may include direct deposition to the water surface, coastal
upwelling from oceanic waters, and transport from watersheds. Freshwater inflows to
estuaries may contain N from agriculture, urban, wastewater and atmospheric deposition
sources. Atmospheric deposition constitutes less than half of the total N supply in most,
but not all, estuaries (Table 7-6). Both point sources and nonpoint sources (including
runoff, as well as atmospheric deposition) have been identified as targets for mitigation of
N loading in coastal areas. The importance of atmospheric deposition as a cause of
estuarine eutrophication is determined by the relative contribution of the atmospheric
versus nonatmospheric sources of N input. Seawater contains high concentrations of
SO42 . so atmospheric inputs of S are unlikely to contribute substantially to
biogeochemistry or biological effects in coastal areas.
1.2.2.3 Critical Loads Concept and General Approaches
The following section provides a discussion of important concepts regarding critical
loads. The definition of a critical load is, "a quantitative estimate of an exposure to one or
more pollutants below which significant harmful effects on specified sensitive elements
of the environment do not occur according to present knowledge" (Nilsson and Grcnnfclt.
1988). It is intended as background material to support better understanding of the critical
load calculations presented throughout the ISA. The main concepts presented here
include CLs as an organizing principle, CL heterogeneity across the landscape, more than
one critical load for a given location, the pros and cons of methods used to calculate
critical loads (e.g., empirical, steady-state, and dynamic), and a comparison of CLs versus
target loads. Uncertainty in calculating CLs is discussed in Chapter 1.13.
June 2018
12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Throughout this ISA, the CL concept is used as an organizing principle to relate
atmospheric deposition to ecological endpoints that indicate impairment. The generally
accepted definition of a CL is by Nilsson and Grennfelt (1988) as: "The quantitative
estimate of an exposure to one or more pollutants below which significant harmful effects
on specified sensitive elements of the environment do not occur according to present
knowledge." The development of a quantitative critical load estimate requires a number
of steps. An illustrative example of the eight general steps is shown in Chapter 1,
Figure 1-3.
1) Disturbance
Acidification
Eutrophication
2) Receptor
Forest
Lake
Grassland
Lake
3) Biological
indicator
Sugar
Maple
Norway
Spruce
Brook trout
Fish species
richness
Species
diversity
Primary
productivity
4) Critical
biological
response
Failure to
reproduce
Seedling
death
Presence
absence
Species
loss
Species
loss
Excess
productivity
5) Chemical
indicator
Soil % Base
Saturation
Soil Ca/AI
ratio
Lakewater
ANC
Lakewater
ANC
Soil C/N
ratio
Lakewater
no3
6) Critical
chemical
limit
10%
1.0
0 peq/L
50 peq/L
20
10 peq/L
7) Atmospheric
pollutant
SO,, NO,,
nh4
S04, no3,
nh4
so4, no3,
nh4
S04, no3,
nh4
no3, nh4
no3, nh4
8) Critical
pollutant load
???
???
???
???
???
???
Al = aluminum; ANC = acid-neutralizing capacity; C = carbon; Ca = calcium; L = liter; peq = microequivalents; N = nitrogen;
NH4 = ammonium; N03 = nitrate; S04 = sulfate.
Source: U.S. EPA i'2008a).
Figure 1-3 An example of the matrix of information considered in defining
and calculating critical loads (see discussion in text). Note that
multiple alternative biological indicators, critical biological
responses, chemical indicators, and critical chemical limits could
be used.
It is important to recognize that there is no single "definitive" critical load for an
ecological effect. Critical loads estimates reflect the current state of knowledge and the
selected limits, indicators, and responses. Changes in scientific understanding may
include, for example, new dose-response relationships, better resource maps and
June 2018
13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
inventories, larger survey data sets, continuing time-series monitoring, and improved
numerical models.
Calculating multiple critical loads for a given pollutant at a single location is not
uncommon due to the nested sequence of disturbances, receptors, and biological
indicators considered for a given pollutant. Multiple critical load values may also arise
from an inability to agree on a single definition of "harm." Calculation of critical loads
for multiple definitions of "harm" may be deemed useful in subsequent discussions of the
analysis and in the decision-making steps that may follow critical load calculation.
The heterogeneity of all natural environments can affect responsiveness of ecosystems to
deposition load. As an example, the high spatial variability of soils almost guarantees that
for any reasonably sized soil-based "receptor" that might be defined in a CL analysis,
there will be a continuum of CL values for any indicator chosen. Although the range of
this continuum of values might be narrow, there is nevertheless an a priori expectation in
any critical load analysis that multiple values (or a range of values) will result from the
analysis. Given the heterogeneity of ecosystems affected by N and S deposition,
published critical load values for locations in the U.S. vary depending on both biological
and physical factors.
The three approaches to developing CLs (i.e., empirical observation, steady-state
modeling, and dynamic modeling) each have strengths and limitations. It is suggested
that the combined approach of calculating CLs from biogeochemical simulation models
in conjunction with empirical analyses is the most effective way to characterize the
effects of deposition to a given environment (Tenn et al.. 2015).
An important advantage of empirical CLs is that they are based on measured (vs.
modeled) changes in ecological variables in response to N inputs. Consequently, the links
between N deposition and the measured response variable are direct; full process-level
knowledge is not required. Empirical CLs are important for validating CL values
determined with models (Tenn et al.. 2015).
Fenn et al. (2015) discuss that the advantages of models, "are that ecosystem responses to
alternative scenarios can be tested. These might include changes in atmospheric
deposition, disturbance or climatic conditions, and responses to silvicultural treatments,
grazing, fire, and other disturbances. Simulation modeling allows temporal aspects of
ecosystem response in relation to CLs and CL exceedances to be evaluated, including
evaluation of historical and future conditions."
Among the types of modeled CLs, two key ways that steady-state and dynamic modeling
differ is by how they assume ecosystem equilibrium and the required amount of input
data needed to parameterize the model. Steady-state models assume that the ecosystem is
14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
in equilibrium with the critical load of deposition; therefore, the long-term sustainable
deposition is indicated. This is the relevant information needed to provide protection
from deposition in perpetuity as the system comes into equilibrium with the pollutant
critical load. In the U.S., few (if any) ecosystems qualify as steady-state systems.
Therefore, the assumption of equilibrium in the steady-state model is often false. The
steady-state models give no information concerning the time to achieve the equilibrium
or what may happen to the receptor along the path to equilibrium. The recovery of an
ecosystem based on a critical load from a steady-state model may take several hundred
years. In other words, the assumption that attainment of a deposition values below the
steady-state critical load will result in biological recovery within a specified time period
may not be valid. Dynamic models calculate time-dependent critical loads and therefore
do not assume an ecosystem that is in equilibrium. The time-dependent calculation is
relevant information to provide protection from damage by the pollutant within a specific
time frame. As a general rule, the shorter the time frame selected, the lower the critical
load.
Data requirements for steady-state models tend to be much lower than for dynamic
models. Therefore, the data required to conduct dynamic modeling are not available for
as many places as the data required to conduct steady-state modeling. The few
national-scale modeling efforts for both terrestrial and aquatic acidification are both done
with steady-state models for this reason.
The results of all three CL approaches are difficult to extrapolate across geographic
space. Spatially, variation in biological and biogeochemical processes imposed by
climate, geology, biota, and other environmental factors may alter the
deposition-response relationship. Empirical CLs may only be applied with confidence to
sites with highly similar biotic and environmental conditions (Pardo ct al.. 201 la). This is
particularly problematic in areas where deposition has received sparse research
attention—as is sometimes the case for CLs of N deposition related to N-driven
eutrophication (Appendix 6.4). Models may be run at different locations; however, the
data needed to parameterize them is not always available.
The traditional critical load has been defined fNilsson and Grennfelt. 1988) as: "The
quantitative estimate of an exposure to one or more pollutants below which significant
harmful effects on specified sensitive elements of the environment do not occur
according to present knowledge." Fenn et al. (2011b) have defined the "target load" as
follows: "The acceptable pollution load that is agreed upon by policy makers or land
managers. The target load is set below the critical load to provide a reasonable margin of
safety, but could be set higher than the critical load at least temporarily." Target loads are
15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
selected based on the level of ecosystem protection desired, economic considerations, and
stakeholder input at a given location.
1.2.2.4 The Importance of Biodiversity
There are causal relationships between N and/or S and biodiversity loss in terrestrial,
freshwater, wetland, and estuarine ecosystems in the U.S. (Chapter 1, Table 1-1). What
does it mean to lose biodiversity? Biodiversity loss not only represents the extirpation of
unique living species; several decades of research link biodiversity to ecosystem function
and ecosystem services in a wide variety of natural systems (Hooper et al.. 2012;
Balvanera et al.. 2006; Tilman. 2000). Numerous studies demonstrate that the number
and diversity of organisms in a system control the abundance of habitat for other species,
the biogeochemical cycling of nutrients and carbon, and the efficiency at which biotic
systems are able to transform limited resources into biomass (Cardinale et al.. 2011).
Among plant communities, higher biodiversity leads to higher overall plant productivity
and greater retention of soil nutrients (Reich et al.. 2012; Tilman. 2000). In multitrophic
systems, higher prey diversity leads to both higher predator growth rates and a smaller
impact of predation on prey abundance (Duffy et al.. 2007). Positive impacts of
biodiversity on ecosystem services have been documented in forests (Gamfeldt et al..
2013; Zhang et al.. 2012b). grasslands (Tilman et al.. 2012). arid and semiarid
ecosystems (Maestre et al.. 2012). and marine systems (Gamfeldt et al.. 2015; Worm et
al.. 2006) and include effects such as greater carbon storage, fruit production, wood
production, and nutrient cycling. In marine ecosystems, biodiversity loss has been linked
to increased rates of resource collapse and exponential decreases in water quality through
metrics such as higher numbers of beach closures and harmful algal blooms [HABs;
(Worm et al.. 2006)1. Notably, HABs are linked to increased disease prevalence among
humans, domestic animals/pets, and aquatic organisms (Johnson et al.. 2010). In addition
to the relationship between HABs and disease, there is now empirical evidence from
many ecosystems of a broader link between declines in biodiversity and increased
transmission and severity of disease (Johnson et al.. 2015b) caused by plant, wildlife, and
human pathogens. As a whole, these decades of research have produced an overwhelming
body of evidence indicating that the loss of biodiversity risks a deterioration of the
ecosystem goods and services that humanity depends on (Gamfeldt et al.. 2015; Cardinale
et al.. 2012).
One of the most important consensus observations in biodiversity research is that
ecosystem processes are more stable (have less temporal variability) at higher levels of
diversity (Cardinale et al.. 2012; McCann. 2000; Naeem and Li. 1997; Tilman and
Downing. 1994). This stability occurs because species respond differently to
June 2018
16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
environmental variation. In diverse communities, is it more likely that declines in the
growth of one species caused by an environmental change will provide more resources
for competing species (Cardinalc et al.. 2012; Tilman. 2000). This property was predicted
by economists and is similar to how more diversified investment portfolios provide
enhanced stability under fluctuating market conditions (Doak et al.. 1998; Tilman et al..
1998). Notably, there is also consensus that the impact of biodiversity on ecosystem
processes is nonlinear, wherein declines in ecosystem processes accelerate as the number
of species in a system declines (Cardinalc et al.. 2012). Accelerating ecosystem service
declines in response to species loss may be due to multifunctionality, which suggests that
different ecosystem functions require the presence of different sets of species (Isbell et
al.. 2015; Reich et al.. 2012; Zavaleta et al.. 2010). The increased stability of diverse
ecosystems makes these systems less vulnerable to environmental change or collapse
caused by external forces such as drought or human disturbance (Isbell et al.. 2015;
Tilman etal.. 2012; Isbell etal.. 2011; Worm et al.. 2006). For example, coastal systems
with higher species diversity had lower rates of fishery collapse and extinction for
commercially important fish and invertebrate species, and large marine ecosystems with
higher fish diversity recovered more quickly from collapse (Worm et al.. 2006). Thus,
there is strong evidence that high biodiversity helps sustain ecosystem services and also
makes these ecosystem services more resilient to environmental change.
1.2.2.5 Reduced versus Oxidized Nitrogen Effects across Ecosystems
Individual biochemical and geochemical processes involve specific chemical forms of N,
suggesting that there may be consequences in many ecosystems from the ongoing trend
of decreasing NOy deposition and increasing NHx deposition in many parts of the U.S.
(Chapter 1.3). The largest body of evidence that the effects of reduced versus oxidized N
may have different consequences for ecological structure and function is for estuaries
where the form ofN delivered to some coastal areas of the U.S. is shifting from primarily
N03 to an increase in reduced forms of N. Although unlikely to be attributed solely to
atmospheric sources due to the large contribution of N from wastewater, agriculture, and
other sources, inputs of NH3/NH44" selectively favor specific phytoplankton functional
groups (e.g., cyanobacteria, dinoflagellates) including harmful species (Figure 10-7).
Shifts in phytoplankton community composition to species that respond strongly to
reduced N have been observed in some coastal regions (Appendix 10.3.2). Growth of
some species of phytoplankton (Appendix 10.2.2) and macroalgae (seaweed;
Appendix 10.2.3) appear to be related to the form of N. There is also increasing evidence
in freshwater systems for the importance ofN in harmful algal blooms (HABs), and
several studies have shown that the form ofN influences freshwater algal species
June 2018
17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
composition (Appendix 9.2.6.1). In terrestrial systems, redox status of inorganic N seems
to have little influence on the biological responses to N deposition (Appendix 4.3.12V
Because some soil biogeochemical processes involve specific chemical forms of N
(e.g., denitrification, ammonium toxicity), there is the potential that biological responses
to N deposition (or N addition) could depend on whether the dominant form of deposited
N is oxidized (NOy) or reduced (NHx). Different responses to individual forms of N have
been observed for some soil biogeochemical processes (Table 4-13) and terrestrial
biological responses (Table 6-1). Moreover, a number of individual studies have
observed differential effects of NH4+ versus NOs additions on plant community diversity
[e.g., (Dias et al.. 2014; Kleijn et al.. 2008)1. In general, however, meta-analyses in the
literature have tended to find no difference in the effects of individual forms ofN on
terrestrial biological endpoints like plant productivity or microbial biomass (Table 6-1).
This result suggests that terrestrial community diversity is also generally not affected,
possibly because plant uptake of N is mediated by soil biogeochemical cycles that often
rapidly transform N between oxidized and reduced forms.
Evidence of wetland responses to different chemical forms of N come primarily from N
addition experiments conducted outside of the U.S. In European bogs and fens, both
forms of N addition decreased ecosystem N retention, but oxidized N addition caused
DON leaching, while reduced N caused DIN as well as cation leaching
(Appendix 11.3.1.6). Reduced N caused greater physiological stress or injury than
equivalent loads of oxidized N in moss species (Appendix 11.4.5 and Appendix 11.5.5).
1.2.2.6 Scientific Advancements of the Aquatic Acidification Index (AAI)
The 2017 IRP (U.S. EPA. 2017b) described the Aquatic Acidification Index (AAI) as a
novel approach for a multipollutant standard intended to address deposition-related
effects that was developed in the 2011 NOxSOx Policy Assessment (U.S. EPA. 2011a).
Scientifically, the AAI represented an advancement in ecological methodology to
(1) calculate critical loads for aquatic acidification on a national scale, when previously
critical loads had been calculated on the spatial scale of a watershed and (2) provide a
uniform level of ecological protection at the national scale. These advancements were
accomplished by first aggregating critical loads calculated for the same chemical limit
within a defined spatial region. Next, the distribution of the "population" of critical load
values was evaluated, and the percentage of water bodies to protect was selected. The
AAI also presented novel advancement in atmospheric sciences, including (1) using
transference ratios to relate atmospheric concentrations of criteria pollutants to deposition
levels and (2) allowing quantification of criteria pollutants (NOy and SOx) and
June 2018
18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
noncriteria pollutant contributions to total acidifying deposition. As a scientific
publication, the AAI is documented in Scheffe et al. (2014). The AAI was originally
developed in the 2011 NOxSOx Policy Assessment (U.S. EPA. 2011a). and the equation
is described with terms that traditionally define aNAAQS [the indicator,1 averaging
time,2 form,3and level4—further described in the 2017 IRP (U.S. EPA. 2017b).
Key scientific aspects of the AAI equation, as the form of a potential standard, are
described in the following excerpt from 2017 IRP (U.S. EPA. 2017b):
"The AAI, as described in the PA (U.S. EPA. 2011a). was constructed
from steady-state ecosystem modeling, and included atmospheric
transference ratios and deposition of reduced forms of nitrogen
(ammonia gas and ammonium ion, expressed as NHx). These
nonoxidized forms of nitrogen were included since ecosystems respond
to total nitrogen deposition, whether from oxidized or reduced forms.
More specifically, the AAI equation was defined in terms of four
ecological and atmospheric factors and the ambient air indicators NOy
and SOx:
AAI = F1-F2- F3[N0y] - F4[SOx]
where F15 represents the ecosystems natural ability to provide
acid-neutralizing capacity (e.g., geology, plant uptake of nitrogen
deposition) and other processes; F26 represents acidifying deposition
associated with reduced forms of nitrogen, NHx; and F31 and 1''4" are the
1	The "indicator" of a standard defines the chemical species or mixture that is measured in determining whether an
area attains the standard.
2	The "averaging time" defines the time period over which ambient measurements are averaged (e.g., 1-hour, 8-hour,
24-hour, annual).
3	The "form" of a standard defines the air quality statistic that is compared to the level of the standard in determining
whether an area attains the standard.
4	The "level" defines the allowable concentration of the criteria pollutant in the ambient air.
5	F1 is defined as: ANC\im + CLJQr, w ith. 1 AY r|lm representing a target ANC level. With regard to the PA
developed distributions of calculated critical loads for a specific ecoregion; in setting an AAI-based standard, a
percentile would need to be specified to reference the value of ('I,, to be used in the AAI equation [(U.S. EPA.
2011a). p. 7- 37], The PA described the percentile as an aspect of the form for the standard [(U.S. EPA. 2011a).
section 7.7],
6	F2 is defined as: NHX/(A. where NHX is the deposition divided by O, [(U.S. EPA. 2011a). p. 7-37],
7	F3 is defined as: 7N0y/6>i. where 7NOy is the transference ratio that converts deposition of NOy to ambient air
concentrations of NOy [(U.S. EPA. 2011a). p. 7-37],
8	F4 is defined as: TSO-JQ,. where '/SOx is the transference ratio that converts deposition of SOx to ambient air
concentrations of SOx [(U.S. EPA. 2011a). p. 7-37],
June 2018
19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
transference ratios that convert concentrations of NOy and SOx to related
deposition of nitrogen and sulfur IYU.S. EPA. 2011a). Section 7.7]."
There are several other key scientific considerations included in the AAI that were
discussed in the 2011 NOxSOx Policy Assessment (U.S. EPA. 2011a):
•	Spatial heterogeneity of factors in the AAI equation: The value of factors in the
AAI equation vary across the U.S. Factors could be calculated for a spatial
boundary based on ecologically similar landscape (e.g., Omernick ecoregion).
•	Temporal heterogeneity: There is a relatively high degree of interannual
variability expected in the AAI because it is so strongly influenced by the amount
and pattern of precipitation that occurs within a region from year to year, therefore
averaging calculated annual AAI values over 3 to 5 years would provide
reasonable stability.
•	Level: With regard to a level for the AAI, the 2011 NOxSOx Policy Assessment
(U.S. EPA. 201 la) concluded consideration should be given to a level within the
range of 20 to 75 |icq/L. noting that a target ANC value of 20 |icq/L would be a
reasonable lower end of this range, so as to protect against chronic
acidification-related adverse impacts on fish populations which have been
characterized as severe at ANC values below this level.
1.2.3 Changes: New Evidence and Causal Determinations
Since the 2008 ISA, several conceptual changes have occurred in our understanding of
the sources of N deposition and in the relationship between atmospheric concentration
and deposition. (Chapter 1.3 and Appendix 2). Models of N deposition rely on accurate
emissions data. Since the 2008 ISA, NOx emissions have been decreasing but NHx
emissions increasing. As a result, total reactive N emissions (NOx + NHx) have not only
remained steady, but their uncertainty has increased. This is because emissions estimates
that have the lowest levels of uncertainty are from stationary and mobile sources (the
main sources of NOx), and higher levels of uncertainty are associated with agricultural
emissions (the main source of NHx).
A better understanding of the relationship between atmospheric concentration and
deposition has resulted from advances in understanding bidirectional exchange of NH3
and NOx chemistry within canopies. These advances have led to the first efforts to
provide a detailed characterization of N and S deposition on a national scale, by using
both measured and modeled values with the goal of providing estimates of total sulfur
and nitrogen deposition across the U.S.
New evidence since the 2008 ISA increases in the weight of evidence for ecological
effects, confirming concepts previously identified and improving quantification of dose
(deposition)-response relationships, particularly for N deposition. The ecological effects
June 2018
20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
are described by the causality determinations. There are 20 causality statements in this
ISA (Chapter 1, Table 1 -1). Fourteen are causal relationships repeated from the 2008 ISA
or modified from the 2008 ISA to include specific endpoints. One is a likely causal
relationship repeated from the 2009 PM ISA. Five are new endpoint categories not
evaluated in the 2008 ISA: three with causal relationships, one with a likely causal
relationship, and one suggestive of a causal relationship. Chapter 1, Table 1-3 shows that
N and S deposition cause alteration of (1) biogeochemical components of soil and water
chemistry and (2) multiple levels of biological organization ranging from physiological
processes to shifts in biodiversity and ecological function (Chapter 1, Figure 1-4).
The current NO2 and SO2 secondary NAAQS are set to protect against direct damage to
vegetation by exposure to gas-phase oxides of nitrogen and oxides of sulfur. Research
continues to support causal relationships between SO2, NO2, NO, peroxyacetyl nitrate
(PAN), HNO3, and injury to vegetation (Chapter 1, Table 1-1). but research that tests
plant response to the lower exposure levels that represent current atmospheric NOy and
SOx concentrations is limited. Therefore, little evidence is available to inform whether
current monitored concentrations of gas-phase NOy and SOx are high enough to injure
vegetation.
It is clear that NOy, SOx, and PM contribute to total N and S deposition, which alters the
biogeochemistry and the physiology of organisms, resulting in harmful declines in
biodiversity. Decreases in biodiversity mean that some species become relatively less
abundant and may be locally extirpated. The current period in Earth's history is the
Anthropocene. In addition to a spike in soil radiocarbon from nuclear bomb testing
(Turncv et al.. 2018). a defining attribute of the Anthropocene is global human-driven
mass extinctions of many species. The biodiversity loss reported in this assessment
contributes to the Anthropocene loss of biodiversity (Rockstrom et al.. 2009). In addition
to the loss of unique living species, the decline in total biodiversity is harmful because
biodiversity is an important determinant of the stability of ecosystems and the ability of
ecosystems to provide services to humanity (see more on biodiversity in Chapter 1.2.2.4).
21
DRAFT: Do Not Cite or Quote

-------
Table 1-1 Causal determinations for relationships between criteria pollutants
and ecological effects from the 2008 NOx/SOx ISA or the 2009 PM
ISA, for other effects of particulate matter (PM), and the current draft
Integrated Science Assessment.

Causal Determination
Effect Category
2008 NOx/SOx ISA
Current Draft ISA
Gas-phase direct phototoxic effects
Gas-phase SO2 and injury to vegetation
Causal relationship
Causal relationship
Appendix 3.5.1


Gas-phase NO, NO2, and PAN and injury to vegetation
Causal relationship
Causal relationship
Appendix 3.5.2


Gas-phase HNO3 and injury to vegetation3
Causal relationship
Causal relationship
Appendix 3.5.3


N and acidifying deposition to terrestrial ecosystems
N and S deposition and alteration of soil biogeochemistry
in terrestrial ecosystems'5
Causal relationship
Causal relationship
Appendix 4.1


N deposition and the alteration of the physiology and
growth of terrestrial organisms and the productivity of
terrestrial ecosystems0
Not included
Causal relationship
Appendix 6.6.1


N deposition and the alteration of species richness,
community composition, and biodiversity in terrestrial
ecosystems0
Causal relationship
Causal relationship
Appendix 6.6.2


Acidifying N and S deposition and the alteration of the
physiology and growth of terrestrial organisms and the
productivity of terrestrial ecosystemsd
Not included
Causal relationship
Appendix 5.7.1


Acidifying N and S deposition and the alteration of
species richness, community composition, and
biodiversity in terrestrial ecosystemsd
Causal relationship
Causal relationship
Appendix 5.7.2


N and acidifying deposition to freshwater ecosystems
N and S deposition and alteration of freshwater
biogeochemistrye
Causal relationship
Causal relationship
Appendix 7.1.7


June 2018
22
DRAFT: Do Not Cite or Quote

-------
Table 1-1 (Continued): Causal determinations for relationships between criteria
pollutants and ecological effects from the 2008 NOx/SOx
ISA or the 2009 PM ISA, for other effects of particulate
matter (PM), and the current draft Integrated Science
Assessment.

Causal Determination
Effect Category
2008 NOx/SOx ISA
Current Draft ISA
Acidifying N and S deposition and changes in biota,
including physiological impairment and alteration of
species richness, community composition, and
biodiversity in freshwater ecosystems'
Causal relationship
Causal relationship
Appendix 8.6


N deposition and changes in biota, including altered
growth and productivity, species richness, community
composition, and biodiversity due to N enrichment in
freshwater ecosystems9
Causal relationship
Causal relationship
Appendix 9.6


N deposition to estuarine ecosystems
N deposition and alteration of biogeochemistry in
estuarine and near-coastal marine systems
Causal relationship
Causal relationship
Appendix 7.2.10


N deposition and increased nutrient-enhanced coastal
acidification
Not included
Likely to be a causal
relationship
Appendix 7.2.10


N deposition and changes in biota, including altered
growth, total primary production, total algal community
biomass, species richness, community composition, and
biodiversity due to N enrichment in estuarine
environments11
Causal relationship
Causal relationship
Appendix 10.7.1


N deposition and changes in biota, including altered
physiology, species richness, community composition,
and biodiversity due to nutrient-enhanced coastal
acidification
Not included
Suggestive of, but not
sufficient to infer, a causal
relationship
Appendix 10.7.2


N deposition to wetland ecosystems
N deposition and the alteration of biogeochemical cycling
in wetlands
Causal relationship
Causal relationship
Appendix 11.10


N deposition and the alteration of growth and productivity,
species physiology, species richness, community
composition, and biodiversity in wetlands
Causal relationship
Causal relationship
Appendix 11.10


June 2018
23
DRAFT: Do Not Cite or Quote

-------
Table 1-1 (Continued): Causal determinations for relationships between criteria
pollutants and ecological effects from the 2008 NOx/SOx
ISA or the 2009 PM ISA, for other effects of particulate
matter (PM), and the current draft Integrated Science
Assessment.

Causal Determination
Effect Category
2008 NOx/SOx ISA
Current Draft ISA
S deposition to wetland and freshwater ecosystems
S deposition and the alteration of mercury methylation in
surface water, sediment, and soils in wetland and
freshwater ecosystems'
Appendix 12.7
Causal relationship
Causal relationship
S deposition and changes in biota due to sulfide
phytotoxicity, including alteration of growth and
productivity, species physiology, species richness,
community composition, and biodiversity in wetland and
freshwater ecosystems
Appendix 12.7
Not included
Causal relationship

2009 PM ISA
Current Draft ISA
Other ecological effects of PM
PM and a variety of effects on individual organisms and
ecosystems
Appendix 15.7
Likely to be a causal
relationship
Likely to be a causal
relationship
C = carbon; Hg = mercury; HN03 = nitric acid; ISA = Integrated Science Assessment; N = nitrogen; NO = nitric oxide;
N02 = nitrogen dioxide; PAN = peroxyacetyl nitrate; S = sulfur; S02 = sulfur dioxide.
aThe 2008 ISA causality statements for gas-phase HN03 was phrased as, "changes in vegetation."
bThe 2008 ISA included two causality statements for terrestrial biogeochemistry which were phrased as, "relationship between
acidifying deposition and changes in biogeochemistry" and "relationship between N deposition and the alteration of
biogeochemical cycling of N."
The 2008 ISA causality statement for biological effects of N enrichment in terrestrial ecosystems was phrased as, "relationship
between N deposition and the alteration of species richness, species composition and biodiversity."
dThe 2008 ISA causality statement for biological effects of acidifying deposition in terrestrial ecosystems was phrased as,
"relationship between acidifying deposition and changes in terrestrial biota."
eThe 2008 ISA included three causality statements for freshwater biogeochemistry phrased as, "relationship between acidifying
deposition and changes in biogeochemistry related to aquatic ecosystems", "relationship between N deposition and the alteration
of biogeochemical cycling of N", and "relationship between N deposition and the alteration of biogeochemical cycling of C."
'The 2008 ISA causality statement for biological effects of acidifying deposition in freshwater ecosystems was phrased as,
"relationship between acidifying deposition and changes in aquatic biota."
9The 2008 ISA causality statement for biological effects of N deposition in freshwater ecosystems was phrased as, "relationship
between N deposition and the alteration of species richness, species composition and biodiversity in freshwater aquatic
ecosystems."
hThe 2008 ISA causality statement for biological effects of N deposition to estuaries was phrased as, "relationship between N
deposition and the alteration of species richness, species composition and biodiversity in estuarine ecosystems."
'The 2008 ISA causality statement for biological effects of S deposition effects on ecosystems was phrased as, "relationship
between S deposition and increased methylation of Hg, in aquatic environments where the value of other factors is within
adequate range for methylation."
June 2018
24
DRAFT: Do Not Cite or Quote

-------
2018 NOx SOx PM Integrated Science Assessment for ECOLOGICAL EFFECTS*
Indicator
Gases + Nitrogen deposition Sulfur deposition Nitrogen and Sulfur
deposition
Class of Pollutant Effect
D- Eutrophication
N-enrichment/Eutrophication driven Sulfide Toxicity Mercury Methylation Acidification
Ph"t0t0XIC acidification
Scale of Ecological Response
Population
Geochemistry Individual 	 Community Ecosystem
Individual
Ecosystem
Productivity
Biodiversity
Growth rate
Physiological
alteration, stress
or injury
Soil or sediment
chemistry
Surface water
chemistry
Terrestrial Terrestrial Wetland Fresh Estuary Estuary Wetland Fresh Wetland Fresh Terrestrial Fresh
Causality framework



U
Causal

Likely causal

Suggestive

Inadequate

Not likely

Not evaluated in causal framework
LI
*	A causal relationship is likely to exist between deposition of PM and a variety of effects on individual organisms and ecosystems, based
on information from the previous review and limited new findings in this review
*	Includes: NO, N02, HN03, S02 and PAN
Figure 1-4 Causal relationships between the criteria pollutants and ecological effects.
June 2018
25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Since the 2008 ISA, the largest increase in ecological evidence is for terrestrial N driven
eutrophication effects (Chapter 1.5.1. Chapter 1.5.2. Appendix 4. and Appendix 6V This
new research confirms the causal relationship between N deposition and ecological
effects documented in the 2008 ISA and improves our understanding of the mechanistic
links that inform causal determinations between N deposition, biogeochemistry, and biota
in terrestrial ecosystems (Chapter 1, Table 1 -1). Further, there is now stronger empirical
evidence from across most regions of the U.S. to quantify the levels of N deposition
(empirical CLs) that cause biodiversity declines of lichens and grasses/forbs. There is
new evidence to quantify empirical CLs across much of the U.S. for nitrate leaching, tree
survivorship, and mycorrhizal biodiversity.
New research confirms that N + S deposition causes terrestrial ecosystem acidification, as
documented in the 2008 ISA (Chapter 1, Table 1-1). New evidence to characterize
terrestrial acidification (soil biogeochemistry changes and biological effects) across large
regions of the U.S. is available; in particular, new modeling work has improved
calculation of CLs for soil acidification (Chapter 1.5.3; Appendix 4 and Appendix 5).
New evidence for freshwater acidification CLs builds on several decades of research
documenting freshwater acidification effects on aquatic biota in the U.S. and confirms the
causal relationships determined in the 2008 ISA (Chapter 1, Table 1-1).
The sources of N driven eutrophication of fresh waters, estuaries, and wetlands include
atmospheric N deposition and N from agricultural and other wastewaters. New research
has helped show how these respective sources contribute to total loading. In freshwater
ecosystems where atmospheric deposition is the primary source of N, such as in high
alpine watersheds, new CLs support previous observations of increased algal
productivity, species changes, and reductions in diversity. New evidence also supports
clear links between aqueous sulfur concentrations in aquatic systems and both mercury
methylation and sulfide toxicity; however, quantitatively linking these outcomes to
atmospheric deposition remains a challenge.
Since the 2008 ISA, N enrichment has been recognized as a possible contributing factor
to increasing acidification of coastal environments. Dissolution of atmospheric
anthropogenic CO2 into the ocean has led to long term decreases in pH. With increasing
N inputs to coastal waters, CO2 in the water column is produced from degradation of
excess organic matter from changing land use, as well as respiration of living algae and
seagrasses, which in turn can make the water more acidic (Appendix 10.5). Increasing
acidification of coastal waters, which can be exacerbated by elevated N input, is
projected to alter marine habitat, have a wide range of effects at the population and
community level and affect food web processes. Nutrient-enhanced coastal acidification
26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
has been documented in systems with strong thermal stratification with spatial or
temporal decoupling of production and respiration processes.
1.3 Emissions and Atmospheric Chemistry
The atmospheric chemistry from emission to deposition discussed in this ISA is for the
criteria pollutants NOy, SOx, and PM. In addition to gas-phase chemical species, total N,
S and N + S deposition is a main focus.
A wide variety of N containing compounds (oxidized + reduced, and organic + inorganic)
contribute to wet and dry N deposition (Appendix 2.1). NHx (NHx = NH3 + NH4+)
includes both the PM component NH4+ and NH3. NH3 may account for 19-63% of total
observed inorganic N deposition, rather than NOy or PM; therefore, it is discussed in the
ISA to better understand how the criteria pollutants NOy and the PM component NH4
contribute with NH3 to cause N deposition. In addition, PM impacts discussed in this
document are also mainly focused on N and S containing species, which together usually
make up most of the PM2 5 mass in most areas of the U.S., and have greater and better
understood ecological impacts than other PM components.
Both gaseous and particulate forms of NOy, SOx, and NHx contribute to atmospheric
deposition. The major components of particulate matter in the U.S. are NO3 . SO42 .
NH4+, organic carbon, and elemental carbon. Of these, NO3 . SO42 . and NH44" usually
have a strong influence on acid deposition, and NO3 and NH44", and in some cases
organic nitrogen (organic nitrates and reduced organic N), make a substantial
contribution to N deposition.
Since the 2008 ISA, there have been a number of new developments including:
•	Expansion of ambient monitoring networks to include NH3 and NOy at selected
sites, and comparisons of monitoring methods with research-grade instruments
(Appendix 2.4);
•	Adoption of new methods, such as data-model fusion, to integrate deposition
information across the U.S. (Appendix 2.5);
•	Incorporation of bidirectional exchange into models of dry deposition
(Appendix 2.5.2); and
•	Improvements in techniques using satellite-based measurements and
chemistry-transport model simulations to estimate emissions, concentrations, and
dry deposition of NO2, SO2, and NH3 (Appendix 2.6).
June 2018
27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
1.3.1
Sources and Atmospheric Transformations
Both gaseous and particulate forms of N and S contribute to atmospheric deposition. The
main contributors to acidifying precipitation are H2SO4 and HNO3, which are formed
from precursor emissions of SO2 and NOx (NO + NO2) (Appendix 2.2). Gaseous
emissions ofNFh and SC^are each dominated by a single source: agriculture (fertilizer
application and animal waste) for NH3 and electricity generating units (EGUs) for SO2.
Notably, SO2 emissions from EGUs have been decreasing. NOx emissions have a wider
distribution of sources, with substantial contributions from highway and off-highway
vehicles, lightning, and EGUs. Primary PM2 5 and PM10 emissions are dominated by dust
and fires, but much of the PM2 5 mass in the U.S. is produced by reaction of gas-phase
precursors to form secondary PM2 5, and the majority of the mass is often due to N and S
species produced by secondary PM2 5 formation. Because of this process, a sharp decrease
in SO2 emissions in recent years has led to a corresponding decrease in SO42 and PM2 5
concentrations.
Major components of particulate N and S include NH4+, NO3 . and SO42 . which are
primarily derived from gaseous precursors NH3, NOx, and SO2 (Appendix 2.3). In the
eastern U.S., NO3 and SO42 make up an even greater portion of PM2 5 mass in areas
where PM2 5 mass is the highest. Formation of particulate N and S is described in the
2009 ISA for Particulate Matter (U.S. EPA. 2009a). An understanding of the sources,
chemistry, and atmospheric processes for these gas-phase and PM species provides a
background for understanding acidifying and N deposition.
1.3.2 Measurement and Modeling Techniques
Monitoring networks across the U.S. measure NOy, SOx, and NHx species involved in
deposition (Appendix 2.4.1). The National Acid Deposition Program/National Trends
Network (NADP/NTN) has monitored precipitation chemistry for several decades at
many U.S. sites. The Clean Air Status and Trends Network (CASTNET) has monitored
concentrations of inorganic gas and particulate-phase N and S species since 1990.
Concentrations are combined with measurements of micrometeorology and surface
characteristics to infer dry deposition. Monitoring of NH3 (Appendix 2.5.3) in the
Ammonia Monitoring Network (AmoN), part of the NADP network, was initiated at a
subset of CASTNET sites in 2007. The Interagency Monitoring of Protected Visual
Environments (IMPROVE) network and the Chemical Speciation Network (CSN)
measure PM and PM components including NH4+, NO3 . and SO42 . although these data
are not used to estimate deposition rates (Appendix 2.4.1).
June 2018
28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Atmospheric N deposition rates are calculated from measurements and models. Direct
measurement of NO2 has limited utility for quantifying NOy deposition rates. Because
NOy is composed of diverse chemical species with a wide range of deposition velocities
and compensation points, unmeasured component species of NOy and concentrations of
all NOy species in data-sparse regions must be provided by regional models in
conjunction with satellite data (Appendix 2.4.2).
Estimates of dry deposition (Appendix 2.5.2) over the contiguous U.S. (CONUS) are
inferred by atmospheric models, either regional-scale chemical transport models (CTMs)
or local-scale micrometeorological models, using CASTNET data. These models perform
well in predictions of long-term changes in PM2 5 sulfate, nitrate, and mass, but are both
subject to uncertainties in their treatment of small-scale turbulence, surface interactions,
and in particular, seasonal variability in NO3 deposition, mainly because of uncertainties
in NH3 emissions. Consequently, dry deposition rates (and ratios of wet-to-dry
deposition) continue to be highly uncertain.
1.3.3 Spatial and Temporal Variability in Deposition
Emissions of SO2 and NOx (NO + NO2) have declined dramatically since the passage of
the Clean Air Act Amendments in 1990. Emissions of NOx in the U.S. from highway
vehicles and fuel combustion declined 49% between 1990 and 2013, while nationwide
annual average NO2 concentrations decreased by 48% from 1990 to 2012 ISA (U.S. EPA.
2016f). Total emissions of SO2 decreased by 72% from 1990 to 2011.
Overall deposition of total N (oxidized + reduced N) has not decreased over the past
25 years (Appendix 2.7). Although NOx emissions have declined in the CONUS,
emissions of NH3 have increased in many areas. The large spatial variability in N
deposition is evident in the map (Chapter 1, Figure 1-5) of average (2011 to 2013) annual
dry + wet deposition of NOy and NHx over the CONUS estimated using the TDEP (Total
Deposition) modeling approach (Appendix 2.7). which combines output from the
Community Model for Air Quality (CMAQ) system with wet deposition from the
NADP/NTN (Schwede and Lear. 2014b).
According to TDEP estimates for 2011-2013 (Appendix 2.7). at least one-third of the
CONUS is estimated to receive at least 10 kg N/ha/yr dry + wet deposition, with some
areas receiving more than 15 kg N/ha/yr. It is likely that estimates for the spatial extent of
the areas receiving at least 10 kg N/ha/yr of deposition and the overall amount of N
deposited are too low because reduced organic N species are not routinely monitored or
considered in air quality models such as CMAQ.
June 2018
29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
In general, wet deposition of reduced N exceeds that of oxidized N across the CONUS.
Nationwide, deposition of N occurs mainly by dry deposition of HNO3 and NH3 (with
NH3 dominant), according to estimates based on CASTNET and NADP data and CMAQ
modeling results (Figure 2-4). Hybrid satellite/modeling and CMAQ results indicate that
dry deposition of NO2 is also a nontrivial source of deposited N in many areas. Over the
past 25 years, NADP/NTN data show that wet deposition of inorganic N
(oxidized + reduced) decreased in areas such as the Northeast, but remained constant or
increased in areas such as the central U.S. (see Figure 2-32 in Appendix 2.7). Wet
deposition of total inorganic N has not tracked declines in NOx emissions over the past
25 years, indicating that wet deposition of reduced inorganic N has increased in this
period.
For S deposition, wet deposition tends to dominate over dry deposition in large areas of
the CONUS. Dry deposition of particulate SO42 is only a minor source of S.
Anthropogenic emissions of S and subsequent deposition have declined markedly since
the 1990s, with the most pronounced declines in the eastern U.S. Currently, the highest
values of total (wet + dry) SOx deposition in the U.S. are in parts of the Ohio Valley
region, and range between 15 to 20 kg S/ha/yr (Figure 2-35).
Both N and S deposition contribute to acidification of ecosystems. The pH of rainwater
has increased markedly across the U.S. since 1990, coincident with decreases in the wet
deposition of nitrate and sulfate. However, there are still widespread areas affected by
acidifying precipitation, mainly in the eastern U.S. (see Appendix 2/7). Total acidifying
deposition (wet + dry N + S, expressed as H+ equivalents) fluxes for 2011 to 2013 ranged
from a few hundred H+ equivalents/ha/yr over much of the western U.S. to over 1,500 H+
equivalents/ha/yr in a broad swath encompassing the Midwest and the Mid-Atlantic
regions, and in other isolated hotspots surrounding areas of concentrated industrial or
agricultural activity (Chapter 1, Figure 1-6).
30
DRAFT: Do Not Cite or Quote

-------
Source: CASTNTTT/T.M AQ.'"NT\"/AMCJN/fvE ARCH
-14
r
¦->20
Total deposition of nitrogen 1113
TTSEPA ia'lft/14
Totai M
(kQ-N/ha)
ha = hectare; kg = kilogram; N = nitrogen.
Figure 1 5 Three-year (2011 to 2013) average annual dry + wet deposition of
total oxidized nitrogen and reduced nitrogen species in kilograms
of nitrogen per hectare per year.
June 2018
31
DRAFT: Do Not Cite or Quote

-------
Eq.
Eq. = equivalents; H+ = hydrogen ion; ha = hectare; yr = year.
Source: NCEA using data from TDEP.
Figure 1-6 Total deposition of total oxidized nitrogen, reduced nitrogen, and
oxidized sulfur expressed as H+ equivalents per hectare per year
over the contiguous U.S. 2011-2013.
1
2	Dry deposition rates are a strong function of surface characteristics, which modify the
3	structure of surface layer turbulence and the resistance to uptake by vegetation
4	(Appendix 2.5.2). As a result, spatially aggregated estimates of dry deposition fluxes are
5	subject to sizable uncertainty, in addition to inherent uncertainties in the measurement of
6	species concentrations and in the inference of dry fluxes. Wet fluxes are not directly
7	influenced by surface characteristics (although orography affects transport and
8	precipitation) but are subject to smaller uncertainties in the measurement of rainfall and
9	chemistry.
H+ Equivalents of Total Nitrogen and Sulfur Deposition
HT/ha/yr
June 2018
32
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
1.4
Gas-Phase Direct Phytotoxic Effects
New evidence supports the causal determinations made in the 2008 ISA regarding
gas-phase effects on vegetation, and there are no new causal statements for gas-phase
effects. As in the 2008 ISA, the current ISA concludes that there are causal relationships
between SO2, NO2, NO, PAN, HNO3, and injury to vegetation. This determination is
based on consistent, coherent, and biologically plausible evidence (Appendix 3.2.
Appendix 3.3. and Appendix 3 A, Chapter 1, Table 1-1). The clearest evidence for these
conclusions comes from studies available at the time of the 2008 ISA. There have been
some additional studies since the 2008 ISA. The majority of evidence on the direct
effects of gaseous NOy and SOx comes from controlled exposure studies across many
species of vegetation. The majority of controlled exposure studies over the past several
decades have used concentrations of gas-phase NOy and SOx above current ambient
conditions observed in the U.S. Relevant information is lacking on exposures and effects
reflecting the more recent lower pollutant conditions. Therefore, there is little evidence
available to inform whether current monitored concentrations of gas-phase NOy and SOx
are high enough to injure vegetation.
1.4.1 Sulfur Dioxide
In the 2008 ISA, evidence was sufficient to infer a causal relationship between exposure
to SO2 and injury to vegetation. The current secondary standard for SO2 is a 3-hour
average of 0.50 ppm, which is designed to protect against acute foliar injury in
vegetation. There has been limited research on acute foliar injury since the 1982 PM-SOx
Air Quality Criteria Document (AQCD), and there is no clear evidence of acute foliar
injury below the level of the current standard. The limited new research since 2008 adds
more evidence that SO2 can have acute negative effects on vegetation but does not
change conclusions from 2008 ISA regarding the causal relationship between SO2
exposure and vegetation damage or the SO2 levels producing these effects (see
Appendix 3.1). Consistent with the 2008 ISA, the body of evidence is sufficient to infer
a causal relationship between gas-phase SO2 and injury to vegetation.
Increased SO2 exposure concentrations and longer exposure times are associated with
decreases in plant growth and yield. The 1982 PM-SOx AQCD concluded that more
definitive concentration-response studies were needed before useable exposure metrics
could be identified. However, very few studies of the effects of SO2 on the growth of
vegetation in the U.S. have been conducted since 1982. Recent studies from eastern
Europe indicate recovery of tree growth in response to decreases in SO2 concentrations
since the 1980s and that annual SO2 concentrations of 4 ppb decreased silver fir (Abies
June 2018
33
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
alba) growth. In West Virginia, the growth of eastern red cedar (Juniperus virginiana)
trees increased with declines in SO2 emissions since the 1980s.
1.4.2 Nitrogen Oxide, Nitrogen Dioxide, and Peroxyacetyl Nitrate
In the 2008 ISA, evidence was sufficient to infer a causal relationship between exposure
to NO, NO2, and PAN and injury to vegetation. It is well known that in sufficient
concentrations, NO, NO2, and PAN can have phytotoxic effects on plants through
decreased photosynthesis and induction of visible foliar injury. However, the 1993
Oxides of Nitrogen AQCD concluded that concentrations of NO, NO2, and PAN in the
atmosphere are rarely high enough to have phytotoxic effects on vegetation (U.S. EPA.
1993). and very little new research has been performed at concentrations currently
observed in the U.S. (see Appendix 3.3). Thus, consistent with the previous 2008 ISA,
the body of evidence is sufficient to infer a causal relationship between gas-phase
NO, NO2, and PAN and injury to vegetation.
1.4.3 Nitric Acid
In the 2008 ISA, evidence was sufficient to infer a causal relationship between exposure
to HNO3 and changes to vegetation. The 2008 ISA reported experimental exposure to
HNO3 resulted in damage to the leaf cuticle of pine and oak seedlings, which may
predispose those plants to other stressors such as drought, pathogens, and other air
pollutants. Since the 2008 ISA, (Padgett et al.. 2009b') investigated dry deposition of
HNO3 on the foliage in a fumigation study and confirmed the earlier research. The 2008
ISA also reported several lines of evidence that past and current HNO3 concentrations
may be contributing to the decline in lichen species in the Los Angeles basin. Subsequent
studies conducted in the Los Angeles basin since the 2008 ISA provide further evidence
of the impacts (see Appendix 3.4). These new studies continue to support the causal
findings of the 2008 ISA, such that the body of evidence is sufficient to infer a causal
relationship between gas-phase HNO3 and changes to vegetation.
1.5 Terrestrial Ecosystem Nitrogen Enrichment and Acidification
For terrestrial ecosystems, new evidence reinforces causal findings from the 2008 ISA
and provides the basis for two new causal statements that reflect a more comprehensive
understanding of how N and acidifying deposition alter terrestrial ecosystem biota
(Chapter 1, Table 1-1). In general, N deposition may cause soil N enrichment and
June 2018
34
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
stimulate the growth of opportunistic species. However, in sensitive soils, deposition of N
and/or S can cause soil acidification, which may decrease growth and cause mortality
among sensitive plant species. Atmospheric deposition ofN and S alter the species
composition of terrestrial systems by one of four mechanisms: (1) nutrient enrichment
(eutrophication; Appendix 4 and Appendix 6). (2) acidification (Appendix 4 and
Appendix 5). (3) direct damage (Appendix 3). and (4) secondary effects (e.g., wildfire;
Appendix 6). Ecosystems and communities may be simultaneously affected by one or
more mechanisms depending on the sensitivity of environmental and biological
properties to each mechanism.
Despite the abundance of N in the environment, plants are unable to directly access the
large pools of N contained in the atmosphere as N2 gas and in the soil as large organic
molecules. Consequently, the limited availability of reactive N often constrains biological
activity in terrestrial ecosystems. Thus, N deposition is considered nutrient enrichment
because N additions generally stimulate plant growth and productivity (cumulative
growth of all vegetation within a community), which has been recognized since the
second half of the 19th century. In comparison, the biological influence of acidifying
deposition is less ubiquitous and largely constrained to ecosystems with historically high
rates of deposition and vulnerable because of factors such as geology and climate. While
S is also an essential macronutrient, less S is required for growth than N, and areas
affected by acidifying deposition typically receive S at rates that greatly exceed biotic
demand. Instead, the impact of acidifying deposition stems from the disruptions to
biochemical processes caused by decreased pH and shifts in soil physiochemical
processes that decrease the supply of other essential nutrients (e.g., Ca, Mg) and increase
the mobilization of toxic forms of Al.
Current knowledge of soil biogeochemistry indicates soil N enrichment and soil
acidification occur in sensitive ecosystems across the U.S. at present levels of deposition.
Newly published work indicates decreasing SO2 emissions and S deposition have led to
early signs of recovery from acidification in some northeastern watersheds, but areas in
the Southeast do not show recovery (Appendix 4). There are many well-defined soil
indicators related to the biological effects of acidifying (N + S) deposition. New evidence
uses these indicators to describe the status of ecosystems, either by empirical observation
or models. Soil indicators for acidification are more typically modeled than those for
eutrophication effects. There is an abundance of new information on biogeochemical
pools and processes, including a new conceptual framework for the N saturation of
terrestrial ecosystems.
The enrichment of terrestrial ecosystems by N deposition often increases plant
productivity and causes changes in physiology and growth rates that vary among species.
35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
This has been observed for herbaceous plants and trees across ecoregions. The changing
growth rates transform competitive interactions between species, and as a consequence
lower species diversity is often observed with increasing N deposition within terrestrial
communities. The relationship between N deposition and community composition is
often derived empirically as critical loads. There are many new critical loads available
since the 2008 ISA, including those for lichens, herbaceous plants, and mycorrhizae.
The process of terrestrial acidification has been well understood and documented for
decades. Recent research, since the 2008 ISA, has confirmed and strengthened this
understanding and provided more quantitative information, especially across
regional-scale landscapes. A number of studies have evaluated the relationships between
soil chemistry indicators of acidification and ecosystem biological endpoints (see
Table 5-6). and some models are well established. There have been new advances in the
parameterization of acidification models to U.S. soils since the 2008 ISA (Appendix 4.5)
resulting in better certainty of critical loads. Biological endpoints included in the
evaluations include physiological and community responses of trees and other vegetation
(such as lichens), soil biota, and fauna.
The following section summarizes the main effects of N and S deposition on terrestrial
ecosystems.
1.5.1 Soil Biogeochemistry
In the 2008 ISA, evidence was sufficient to infer causal relationships between
(1) acidifying deposition and changes in terrestrial biogeochemistry and (2) between N
deposition and terrestrial biogeochemical cycling of N. There is new evidence of how
deposition contributes to total loading in ecosystems, as well as new information from
addition, gradient, and time-series studies characterizing how deposition affects soil pools
and processes. Much of the new work focuses on the effects of N deposition, with
relatively little work focusing on S deposition. Soil N enrichment and soil acidification
occur in sensitive ecosystems across the U.S. at present levels of deposition. Decreasing
S emissions have led to early signs of recovery from acidification in some northeastern
watersheds, but areas in the Southeast do not show recovery. Deposition rates of total N
(NOy + NHx) are relatively unchanged across much of the CONUS (Appendix 2.7).
Accordingly, there are no signs of recovery from N enrichment effects. Critical load
determinations have been made at the ecoregion scale for NOs leaching. Critical loads
for biological effects are summarized below (Chapter 1.5.1.2. Chapter 1.5.2.2. and
Chapter 1.5.3.3). The body of evidence is sufficient to infer a causal relationship
June 2018
36
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
between N and S deposition and alteration of soil biogeochemistry in terrestrial
ecosystems, which is consistent with the conclusions of the 2008 ISA.
1.5.1.1 Soil Processes and Indicators
Deposition ofN or N + S alters soil chemistry, which can have cascading effects on
aquatic ecosystems (for effects on aquatic biology and chemistry see
Appendix 7-Appendix 10). Soil acidification is a natural process that can be accelerated
by N or S deposition. Deposition in the forms of HNO3 and H2SO4 can directly acidify
soils; however, deposition of reduced forms of N (e.g., NHx) can also cause soil
acidification by releasing hydrogen ions (H+) during the microbial oxidation of NH4+ to
NC>3~. There are a number of soil geochemical processes associated with acidification
(Chapter 1, Table 1-2). Base cations counterbalance acid anions. Base cations are added
to the soil solution by weathering and atmospheric deposition and are removed by
leaching and biological uptake. Where acidifying deposition rates are high relative to
base cation input, this deposition can deplete exchangeable base cation pools in soils.
There are several useful indicators of soil acidification (Chapter 1, Table 1-2) that have
quantitative relationships to biological responses (Appendix 5).
Table 1-2 Summary of key soil geochemical processes and indicators
associated with eutrophication and acidification.
Endpoint
N Driven
Nutrient
Enrichment Acidification
The Effect of Deposition
Process
N saturation
X
X	New empirical evidence suggests revising the N saturation
concept, specifically it is now observed that NO3" leaching
can occur even if the ecosystem N capacity to retain N has
not yet been saturated.
Soil N accumulation
X
X	New meta-analysis across ecosystem types confirms
inorganic soil NO3" concentration increases with N addition.
New gradient study confirms that N concentration increases
with N deposition. New addition study confirms increased
soil N accumulation. New studies on Soil N accumulation
are summarized in Table 4-3.
June 2018
37
DRAFT: Do Not Cite or Quote

-------
Table 1-2 (Continued): Summary of key soil geochemical processes and
indicators associated with eutrophication and acidification.
N Driven
Nutrient
Endpoint	Enrichment Acidification	The Effect of Deposition
NO3" leaching	X	X	New meta-analysis confirms leaching increases with N
additions. Regional-scale gradient analyses: <8 kg N/ha/yr
onset of leaching; <1 kg N/ha/yr in European forests; in the
NE U.S., 90% retention for sites receiving 7 kg N/ha/yr to
60% retention for sites receiving 11 kg N/ha/yr.
New USFS critical loads for the onset of leaching: 8-10 kg
N/ha/yr in eastern and western U.S., 17 kg N/ha/yr in the
Sierra Nevada and San Bernardino Mountains. New studies
on Soil N accumulation are summarized in Table 4-3.
Some soils (notably in many watersheds in the SE U.S.)
have the capacity to adsorb substantial quantities of S, with
essentially no acidification of drainage water. Nevertheless,
S adsorption capacity is finite, and under continual high S
deposition loading, the adsorptive capacity of soil will
eventually be exceeded.
New studies of 27 watersheds in the SE indicate most will
begin releasing SO42" in the next two decades; NE
watersheds show a net loss of S from soils now in response
to decreased levels of atmospheric S deposition. New
studies on soil S accumulation are summarized in Table 4-4.
SO42" leaching	X	Atmospheric S deposition generally increases leaching of
SO42" to surface waters. The amount of deposition that
causes the onset of leaching varies across the landscape.
New studies on soil SO42" leaching are summarized in
Table 4-4.
Base cation (Ca, Mg, K, Na) release from soil particles to the
soil solution occurs in response to the input of acid anions
(SO42" and NO3") from deposition.
New studies confirm base cation depletion continues to
occur in the Rocky Mountains (threshold 28 kg N/ha/yr) and
in U.K. grasslands, while in a NE forest, 17 yr of N addition
did not cause further depletion. A meta-analysis suggests
cation depletion early after increased deposition of acid
anions, but this depletion tapers off with time. New studies
on base cation leaching and exchange are summarized in
Table 4-5.
Al mobilization	X	The threshold for inorganic Al mobilization from soil is
<15-20% soil base saturation. This is an extremely
important effect of acidifying deposition because inorganic
monomeric Al is toxic to biota (Appendix 5 and Appendix 8).
Inorganic Al is minimally soluble at pH 6.0, but solubility
increases steeply at pH below 5.5.
New studies on Al in soils are summarized in Table 4-6.
S accumulation and	X
adsorption
Base cation leaching	X
and exchange
June 2018
38
DRAFT: Do Not Cite or Quote

-------
Table 1-2 (Continued): Summary of key soil geochemical processes and
indicators associated with eutrophication and acidification.
Endpoint
N Driven
Nutrient
Enrichment
Acidification
The Effect of Deposition
Nitrification
X
X
Nitrification releases 2 mol hydrogen ion (H+) per mol NHV
converted to NO3", acidifying soils. As soil inorganic N
accumulates, net nitrification rates often increase, and NO3"
can leach from the ecosystem.
New N gradient and meta-analysis studies confirm N
addition increases nitrification. New studies on nitrification
are summarized in Table 4-6.
Denitrifi cation
X

Denitrification is the microbial reduction of NO3" to NO2",
NO, the greenhouse gas N2O, and N2, which occurs under
anaerobic conditions. In Europe, soil switched from a source
to a sink after two decades of N deposition exclusion. New
meta-analysis confirms N addition increases denitrification
rates. New studies on denitrification are summarized in
Table 4-6.
DOC leaching
X
X
In recent years, the DOC of many lakes and streams has
risen, with the source likely from the soils in the adjacent
terrestrial watershed. However, the mechanism causing the
observed increase is unclear and may be due to a
combination of soil recovery from acidification, changes in
climate (e.g., temperature and precipitation), and N
deposition among other mechanisms. New studies are
summarized in Table 4-10.
Decomposition
X
X
The addition of N can stimulate the breakdown of labile
compounds that degrade during the initial stages of
decomposition, but added N can suppress the
decomposition of more recalcitrant material. There are new
addition studies and meta-analysis on mechanisms and
response trends.
New studies are summarized in Table 4-8.
Indicator
Soil [N]
X
X
Increases in soil [N] indicate soil N accumulation and the
size of the soil N pool that may be assimilated by organisms
or mobilized via leaching.
Soil C:N ratio
X
X
Decreasing soil C:N linked to changes in decomposition and
increases in nitrification and NO3" leaching.
<20-25 causes increased nitrification and elevated risk of
NO3" leaching in the U.S. and <25-30 for increased NO3"
leaching in Europe.
Soil base saturation

X
Increasing N + S deposition decreases the soil pool of
exchangeable base cations.
<15-20% exchange ion chemistry is dominated by inorganic
Al and mav cause iniurv to veqetation (see Appendix 5).
June 2018
39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
Table 1-2 (Continued): Summary of key soil geochemical processes and
indicators associated with eutrophication and acidification.
N Driven
Nutrient
Endpoint	Enrichment Acidification	The Effect of Deposition
Soil Bc:AI ratio	X	Increasing N + S deposition decreases the soil pool of
exchangeable base cations, often decreasing the Ca:AI
ratio.
Fungi-to-bacteria ratio	X
Ca:AI <1.0 causes physiological stress, decreased growth,
and mortality of sensitive plant species (see Appendix 5).
New indicator: Increasing N deposition decreases the
fungi-to-bacteria ratio and causes a transition from N to C
limitation among soil food webs.
Al = aluminum; Al3+ = aluminum(lll); Be = base cations; C = carbon; Ca = calcium; DOC = dissolved organic carbon; H+ = hydrogen
ion; ha = hectare; K = potassium; kg = kilogram; Mg = magnesium; N = nitrogen; N2 = molecular (atmospheric) nitrogen;
N20 = nitrous oxide; Na = sodium; NE = northeastern; NH4+ = ammonium; NO = nitric oxide; N02" = nitrite; N03" = nitrate;
S = sulfur; SE = southeastern; S042" = sulfate; U.S. = United States; USFS = U.S. Forest Service; U.K. = United Kingdom;
yr = year.
Some of the same processes and indicators associated with acidification are also
associated with the N enrichment of soils in response to the input of exogenous N
(Appendix 4.3). The 2008 ISA documented that the increase in global reactive N (defined
as NOy + NHx + organic N) that occurred over the previous century was largely due to
three main causes: (1) widespread cultivation of crops that promote conversion of N2 gas
to organic N through biological N fixation, (2) fossil fuel combustion converting
atmospheric N2 and fossil N to NOx, and (3) the Haber-Bosch process, which converts
nonreactive N2 to reactive N to sustain food production and some industrial activities
(Galloway et al.. 2003; Galloway and Cowling. 2002).
The 2008 ISA documented that atmospheric deposition of N can increase soil N. Soil N
accumulation is linked to increased N leaching and decreased retention of N. Critical
loads for the onset of elevated NO3 leaching are given in Appendix 4.6.2.2.
The 2008 ISA described the conceptual model of N saturation, which occurs when N
input rates to terrestrial ecosystems exceed the uptake capacity of the soils and biota, and
is indicated by the onset of increased soil N leaching. However, more recent work has
revised the N saturation model in response to observations in which N leaching resulted
from N input rates that are faster than vegetation and soil uptake rates, thus distinguishing
capacity N saturation from kinetic N saturation. Budgets from 83 forested watersheds in
the northeast U.S. show that N retention averages 76% of the incoming atmospheric N
deposition and decreases from 90% retention at 7 kg N/ha/yr of deposition to 60%
retention at 11 kg N/ha/yr of deposition.
June 2018
40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
The 2008 ISA documented that N enrichment is associated with changes in microbially
mediated biogeochemical processes, including nitrification, denitrification, and
decomposition (Appendix 4.3). The addition of N can increase nitrification (the microbial
conversion of NH44" to NOs ). which contributes to soil acidification. Nitrification is often
stimulated in soils with a C:N ratio below approximately 20 to 25, which can be
decreased by N deposition. The NOs created by nitrification may be leached or
denitrified. Denitrification is the microbial reduction of NO3 to NO2 . NO, the
greenhouse gas N2O, and N2, which occurs under anaerobic conditions. Several syntheses
have been published since 2008 evaluating N addition effects on denitrification and
nitrification in terrestrial ecosystems. A new meta-analysis shows N addition
substantially increases denitrification from many types of ecosystems (e.g., coniferous
forest, deciduous forest, tropical forest, wetland, grassland), except heathlands. Among
five chemical forms of N studied, NO3 addition showed the strongest stimulation of N2O
emission. Using data extracted from 206 peer-reviewed papers, a second meta-analysis
observed that the largest changes in the ecosystem N cycle caused by N addition were
increased nitrification (154%), N2O emissions (134%), and denitrification (84%).
About one-half of C fixed annually during photosynthesis by terrestrial vegetation is
allocated to belowground C processes; therefore, it is important to understand how N
affects belowground C to better understand changes in plant physiology, plant growth,
and ecosystem C cycling (Appendix 6). Many studies have shown that changes in
belowground C cycling do not always parallel shifts in aboveground C cycling, making
extrapolation from aboveground responses to belowground processes inappropriate. New
studies published since 2008 (Appendix 4.3.10) have generally found that N addition
increases aboveground litter inputs (+20%), inhibits CO2 loss via microbial respiration
(-8%), and decreases microbial biomass carbon (-20%). Dissolved organic carbon
concentrations increased (+18%), suggesting C leaching loss may increase. The addition
of N increased the C pool size within the soil organic horizon (+17%), attributed to both
increased litter input and decreased decomposition (inferred from the lower microbial
respiration rates).
The effect of N on organic matter decomposition is an active area of research
(Appendix 4.3.9). Decomposition primarily occurs through the leaching of soluble
chemicals, the depolymerization of complex biomolecules by microbial extracellular
enzymes, and the microbial assimilation of nutrients. Bacteria and fungi are the primary
microbial decomposers of organic matter. Both microbial community composition (see
Appendix 6) and microbial enzyme activity can respond to shifts in inorganic N and
substrate availability. Within the literature, litter decay rates have long been correlated
with the ratio of N to C or lignin in litter. Based on these observations, it could be
assumed that added N would stimulate decomposition and the loss of C from soil pools.
41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
However, the stimulatory effects of N on decomposition are limited to the early stages of
mass loss, and the addition of N slows the later stages of decomposition that are
controlled by the degradation of recalcitrant compounds. In general, N additions decrease
respiration from soil heterotrophs, but do not necessarily decrease total soil respiration
because heterotrophic respiration accounts for only a portion of the soil CO2 efflux.
1.5.1.2 National-Scale Sensitivity and Critical Loads
As of the 2008 ISA, the regions of the U.S. with abundant acid-sensitive soils had been
well delineated. These acid-sensitive ecosystems are generally located in mountainous
terrain in the eastern U.S. and are underlain by bedrock resistant to weathering. However,
a similar delineation of the areas sensitive to the eutrophication effects of N had not yet
been completed. There is strong evidence demonstrating that biogeochemical sensitivity
to deposition-driven eutrophication and acidification is the result of historical loading,
geologic/soil conditions (e.g., mineral weathering and S adsorption), and/or natural
sources of N and S loading to the system. There is no single deposition level applicable to
all ecosystems in the U.S. marking the onset of eutrophication or acidification.
Several new publications report the results of field observations and modeling studies on
soil recovery from acidification, specifically in the northeastern U.S., and the lack of
recovery in the southern Appalachian Mountains (Table 4-18). New ecoregion-scale
terrestrial critical loads for NO3 leaching were published in 2011 and have been updated
by more recent published work. However, the most recent national-scale assessment of
soil acidification was published in 2007.
1.5.2 Biological Effects of Terrestrial Nitrogen Enrichment
The enrichment of terrestrial ecosystems by N deposition often increases plant
productivity and causes changes in physiology and growth rates that vary among species.
This combination of effects can alter the composition and decrease diversity of terrestrial
communities by transforming competitive interactions between species and changing the
availability of other essential resources, including light, water, and nutrients. Because N
deposition can cause both eutrophication and acidification and these processes can occur
simultaneously, the relationship between N deposition and community composition has
often been derived empirically. Many of the effects of N deposition are similar across
ecosystems and life forms because N is an essential macronutrient, but the composition
and magnitude of how these effects are expressed within an ecosystem can differ as a
result of biotic and abiotic influences. Consequently, as with the 2008 ISA, we have
June 2018
42
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
grouped the effects of N deposition on physiology and biodiversity by biome (e.g., forest,
tundra, grassland, and arid lands), with further framing by life form (e.g., plants,
microorganisms) and functional groups (e.g., trees, herbaceous plants). In comparison,
the broadest CLs created by the scientific community are at the ecoregion level, in which
spatial boundaries across the landscape are typically defined based on ecological,
climatological, and geological differences.
The 2008 ISA documented consistent evidence that N additions increased plant
productivity across a wide range of terrestrial ecosystems. Since 2008, a large body of
new research on the biological effects of added N in terrestrial ecosystems has been
published from investigations of plant and microbial physiology, long-term
ecosystem-scale N addition experiments, regional and continental-scale monitoring
studies, and syntheses. These studies have been conducted in ecosystems representing
biomes in the U.S., including tundra, grasslands, arid and semiarid lands, and tropical,
temperate, and boreal forests. As a consequence of the breadth of this research, there is a
strong mechanistic and empirical understanding for many of the biological effects of
added N. This body of evidence is sufficient to infer a causal relationship between N
deposition and the alteration of the physiology and growth of terrestrial organisms
and the productivity of terrestrial ecosystems.
The varying effects of N deposition on the growth and physiology of individual species
have consequence for biodiversity. In the 2008 ISA, evidence was sufficient to infer a
causal relationship between N deposition and the alteration of species richness, species
composition, and biodiversity in terrestrial ecosystems. The 2008 ISA documented
consistent evidence of reduced species richness and altered community composition from
N addition studies in the U.S. and N deposition gradient studies in Europe for grassland
plant diversity, forest understory plants, and forest mycorrhizal fungi. There was also
consistent evidence of altered plant and mycorrhizal community composition from N
addition studies in arid and semiarid ecosystems, particularly in southern California.
There was little evidence of changes in forest overstory tree composition. Since the 2008
ISA, new research techniques have been developed to understand community
composition, a larger number of communities have been surveyed, and new regional and
continental-scale studies have made it possible to isolate the influence of N deposition
from other environmental factors. This new research has provided more extensive and
mechanistic evidence, and combined with the findings of the 2008 ISA, this body of
evidence is sufficient to infer a causal relationship between N deposition and the
alteration of species richness, community composition, and biodiversity in
terrestrial ecosystems.
43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
1.5.2.1 Physiology and Biodiversity
At the time of the 2008 ISA, terrestrial ecologists had used meta-analyses to broadly
quantify the effects that N deposition can have on the growth of terrestrial plants,
concluding that N additions stimulate plant productivity by 20-30% in grasslands,
forests, tundra, and wetlands, increase aboveground productivity in herbaceous plant
communities, alter plant tissue chemistry, decrease biomass of mycorrhizal fungi, and
alter litter decomposition (Appendix 6.6.1). Recent research has provided further
coherent and consistent evidence that N additions stimulate plant growth and
productivity, but this research is still dominated by studies of temperate ecosystems and
aboveground plant responses (Figure 6-1 and Figure 6-2).
In the 2008 ISA, the positive plant growth response to N deposition was attributed to
higher rates of photosynthesis. However, evidence for this is mixed: increases in
photosynthesis following N additions have been observed across a variety of plant
functional types, but higher rates of photosynthesis have not been consistently observed
in response to chronic N additions meant to simulate atmospheric deposition. There is
new support for another mechanism that would increase aboveground growth: decreases
in the quantity of C allocated by plants to roots, mycorrhizae, and root exudation. There
was evidence in the 2008 ISA that N additions increase aboveground biomass more than
belowground biomass, raising the shoot-to-root ratio among plants, but evidence is now
more consistent and widespread. Plants invest substantial amounts of C to support
mycorrhizal fungi, but there is evidence this investment declines when N is added to
terrestrial ecosystems. Similarly, there is mounting evidence that plants can increase root
exudation as N availability decreases.
Evidence that biodiversity change can be a consequence of N deposition has accumulated
since 2008 and includes new information for major taxonomic groups, including
herbaceous plants, overstory trees, and two groups of symbionts (lichens and
mycorrhizae). Evidence is now more widespread for decreases in lichen species richness
as the result of N deposition in the U.S. There are direct observations that N deposition in
the U.S. is changing mycorrhizal community composition and altering herbaceous plant
species richness across a broad range of ecosystems, including forests, grasslands, arid
and semiarid ecosystems, and alpine tundra. In addition, based on changes in mortality
and growth rates of dominant tree species, there is also indirect evidence that N
deposition is altering overstory tree community composition.
A substantial body of research linking changes in biodiversity to shifts in N availability
has been developed. Within this body of research, there is evidence that (1) rare species
are particularly vulnerable to loss and (2) organisms with specific traits will have either
positive or negative responses in growth and survival when N is added. Both mechanisms
June 2018
44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
can operate simultaneously and both mechanisms tie the changes in physiology, growth,
and productivity caused by increased N availability to declines in biodiversity.
As noted in Appendix 4. soil microorganisms have important roles in regulating N and C
cycling. There are several mechanisms to alter soil microbial biomass and physiology,
including changes in soil pH, increases in inorganic N availability, shifts in soil food
webs, and changes in the quantity and quality of available C. There were some
observations in the 2008 ISA that added N decreases microbial biomass, but there is now
more evidence that added N generally negatively or neutrally affects microbial biomass C
and microbial biomass N (Table 6-4).
Forests
Forests occur within every U.S. state, but are most abundant in the eastern U.S., montane
and coastal portions of the western U.S., and Alaska. The distribution of forests is bound
by water availability, cold temperatures, and land management. In the 2008 ISA, there
was consistent evidence that N additions stimulated forest productivity, but these
responses varied widely and included both neutral and negative effects of N additions on
tree growth. However, there had been no empirical analyses of how atmospheric N
deposition altered forest productivity in the U.S. at broad scales. The 2008 ISA lacked
information on whether N deposition had any impact on the diversity and composition of
forest overstory trees, but did present evidence for changes in the composition of
herbaceous vegetation, epiphytic lichens, and microbial communities. The addition of
new research since the 2008 ISA provides coherent evidence that N deposition alters the
physiology, growth, and community composition of overstory trees, understory plants,
lichens, mycorrhizal fungi, soil microorganisms, and arthropods.
As of the 2008 ISA, most long-term N addition experiments were located in temperate
forests in the northeastern U.S. or in temperate or boreal forests in Europe. In these
studies, conifer species were less likely than broadleaf species to exhibit positive growth
responses to added N and more frequently exhibited increased mortality and decreased
growth. Since the 2008 ISA, a number of new observations from experiments, forest
inventory studies, model simulations, and data synthesis efforts have been published
quantifying increases in forest net primary productivity (NPP), net ecosystem
productivity (NEP), and ecosystem C storage (Figure 6-3). Overall, evidence is consistent
that N deposition increases forest ecosystem C storage, including specific evidence
indicating that current rates of N deposition in the northeastern U.S. broadly stimulate
aboveground forest productivity.
45
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Many of the observations in the 2008 ISA about how long-term N additions affected
forest mortality and aboveground productivity have been reinforced by more recent
research, including long-term forest inventory data collected from across the eastern U.S.
and in Europe. Growth and mortality responses have apparent links to plant functional
traits; for example, several conifer species common to the northeastern U.S. exhibited
negative growth responses in both long-term N addition experiments and in an analysis of
forest inventory data. Tree species hosting arbuscular mycorrhizal fungi also showed
increased growth in response to N additions and N deposition.
Analyses of forest inventory from the eastern U.S. have not directly assessed changes in
overstory tree composition, but the evidence of species-specific effects on growth and
mortality observed in these studies suggest that changes in community composition are
occurring. These analyses represent an advancement in our understanding from the time
of the 2008 ISA, when the impact of N deposition on the composition of forest overstory
trees was unclear.
In comparison, there is direct evidence that N deposition is altering the composition of
forest understory plant communities. The evidence for altered forest understory plant
communities (also known as herbaceous layer or groundcover vegetation) comes from
both the 2008 ISA and from the more recent literature. Changes in understory plant
communities have been observed in monitoring plots along atmospheric N deposition
gradients in the U.S. and in Europe. In Europe, forest understory plant communities have
shifted toward more nutrient-demanding and shade-tolerant plant species.
Higher rates of aboveground tree growth in response to N deposition might be due to
shifts in C allocation away from belowground processes. Changes in C allocation in
response to additional N have been accompanied by decreases in the abundance of
mycorrhizal fungi and changes in mycorrhizal community composition (Table 6-2.
Table 6-14). Evidence for composition change is particularly abundant in
ectomycorrhizal fungal communities (Table 6-14); there are fewer observations of how
arbuscular mycorrhizal fungal communities change in response to N additions (Table 6-3;
Table 6-16). There are also numerous observations of altered total microbial (including
bacterial) biomass and community composition. For microbial biomass, most studies
identified since 2008 observed either negative or neutral effects ofN additions, consistent
with results of syntheses published before the 2008 ISA (Table 6-4). Changes in soil
microbial community composition were identified along an N deposition gradient in
Europe and in all three N addition studies. The effects of N additions on individual
microbial taxonomic groups (bacteria, archaea, fungi, etc.) have been less consistent
(Table 6-15). Overall, there is evidence that N additions can decrease total microbial
biomass and alter microbial communities in forest soils.
46
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Within soil food webs, soil microorganisms have both direct and indirect links to
arthropods. Because arthropods feed upon both microorganism and litter, they can be
important regulators of decomposition, nutrient cycling, and forest productivity. Several
studies have examined the response of forest arthropod communities to added N,
including a group of studies on insect herbivores conducted in southern California
(Table 6-17). There is coherent evidence that N additions can alter forest arthropod
communities.
Epiphytic lichens have long been recognized as sensitive to air pollution. Although these
organisms often make up a small portion of forest biomass, they have important roles in
hydrologic cycling, nutrient cycling, and as sources of food and habitat for other species.
New research on lichen community composition identified since the 2008 ISA has further
added to the consistent and coherent evidence that lichen communities in the U.S. and
Europe are sensitive to current levels of atmospheric N deposition (Appendix 6.2.6;
Table 6-23). In particular, the U.S. Forest Service's Forest Inventory and Analysis
Program has ample data on the abundance of lichens throughout the U.S., and shifts in
lichen community composition clearly attributable to atmospheric N pollution have been
observed in forests throughout the West Coast, in the Rocky Mountains, and in
southeastern Alaska. Shifts in epiphytic lichen growth or physiology have been observed
along atmospheric N deposition gradients in the highly impacted area of southern
California, but also in more remote locations such as Wyoming and southeastern Alaska,
(Table 6-5). A number of experimental N studies have also created more detailed insight
into changes in lichen physiology processes. Overall, there is widespread evidence from
forests that N deposition can alter the growth, physiology, and biodiversity of trees,
herbaceous plants, lichens, soil microorganisms, and arthropods.
Tundra
Within the U.S., tundra ecosystems are limited to Arctic ecosystems in Alaska and to
relatively isolated, high elevation sites in the West. Although these ecosystems tend to be
remote, the influence of atmospheric N deposition is distinct and there was evidence in
the 2008 ISA indicating that alpine tundra plant communities were sensitive to
atmospheric N deposition. Alpine organisms may be more sensitive to N deposition
because of the unique nature of N cycling in these ecosystems, which tend to have limited
inorganic N availability. Since the 2008 ISA, numerous studies of tundra physiological,
productivity, and community composition responses to added N have been published,
providing further evidence that N deposition alters the growth and physiology of alpine
plant communities, including vascular plants (herbaceous and woody), bryophytes, and
47
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
lichens (Appendix 6.2.4). as well as evidence of altered soil microbial communities
(Table 6-8: Table 6-19V
As in forests, increases in N content in response to additional N are widespread in tundra
plant communities (Table 6-6). Higher tissue N concentrations in response to added N
have been observed in multiple studies for vascular plants, bryophytes, and lichens. The
2008 ISA noted that plant growth and biomass responses tended to be species specific.
Subsequent studies have confirmed this result (Table 6-6). showing varying responses to
added N among ecosystem types, plant functional groups, and species. Whereas vascular
plants tend to show a positive response to added N, both bryophytes and lichens tend to
decrease in biomass and cover (Table 6-5; Table 6-6).
Given the varying effects of N addition on species physiology and growth, the numerous
observations of N addition impacts on species richness, species diversity, and community
composition among vascular plants, bryophytes, and lichens in alpine and Arctic tundra
ecosystems are unsurprising (Appendix 6.3.4; Table 6-18). Within the U.S., these
observations have included effects of N additions on plant community composition in
Colorado and Washington. In northern Europe, decreases in plant species richness along
atmospheric N deposition gradients have been documented. Overall, this new research
has provided further evidence that experimental N additions can alter plant biodiversity in
alpine and Arctic tundra ecosystems and has provided new evidence that current rates of
atmospheric N deposition in Europe are associated with a loss of plant species richness in
these ecosystems.
There are relatively few observations regarding the effect of N additions on total
microbial biomass or the biomass response of individual microbial taxonomic groups in
tundra ecosystems, and these results have also been largely inconsistent. However, new
research has provided evidence that N additions can alter microbial community
composition in alpine tundra ecosystems (Table 6-8; Table 6-19).
Grasslands
Grasslands are most prevalent in the central U.S., yet also are widely distributed across
the U.S. in areas where woody vegetation is excluded by environmental factors or
management. There was widespread evidence at the time of the 2008 ISA that the
growth, physiology, and productivity of grassland plants could be altered by N
deposition. In addition, there were multiple lines of evidence in the 2008 ISA that
grassland plant, mycorrhizal, and microbial communities were sensitive to N inputs.
Combined with subsequent research, the evidence is clear that physiology, growth, and
48
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
community composition of plants, mycorrhizae, soil microorganisms, and arthropods are
sensitive to N inputs in grasslands.
Although NPP can be limited by multiple factors (e.g., water, herbivores, other nutrients)
in all ecosystems, these other limitations tend to be more marked in grasslands than
forests, making it harder to understand and predict the effects of increased N availability.
However, the general response is similar (Appendix 6.2.5): N additions stimulate NPP,
increase foliar N, and increase allocation to aboveground biomass (increased ratio of
shoot:root mass).
Evidence from the U.S. of grassland plant community composition change in the 2008
ISA was based on N addition studies in Mediterranean grasslands in California and
northern prairie ecosystems. However, large-scale assessments of biodiversity across
observed atmospheric N deposition gradients were restricted to Europe. Recent research
provides further evidence that N deposition reduces grassland biodiversity in the U.S. and
Europe (Appendix 6.3.5). Since 2008, there have been direct observations of reduced
species richness along atmospheric N deposition gradients for grasslands in the U.S. and
Europe. These gradient studies have documented an interaction with soil pH, noting that
N deposition causes a greater loss of species richness and a shift in community
composition at sites with lower pH. Together, these findings from deposition gradients in
the U.S. and Europe provide coherent evidence that N deposition causes shifts in plant
community composition and the loss of plant species richness through mechanisms of
both acidification and eutrophication. Experimental studies published since 2008 have
provided more insight into the mechanisms linking changes in plant and microbial
community composition to increased N availability, providing evidence that declines in
species richness increase with time and that competition for resources such as water may
exacerbate the effects of N addition on diversity.
Overall, the additional studies in grassland ecosystems have confirmed that many of the
responses observed in the older N addition studies also occur at present rates of
atmospheric N deposition. These changes include losses in forb species richness (which
make up the majority of grassland biodiversity), greater growth of grass species (which
make up the majority of grassland biomass), changes in reproductive rates, as well as
shifts in mycorrhizal (Table 6-16). soil microbial (Table 6-20). and arthropod
populations. In total, due to the prevalence of N limitation in grasslands and the
dominance by fast-growing species that can shift in abundance rapidly (in contrast to
forest trees), grasslands appear especially sensitive to N input rates comparable to N
deposition across much of the contiguous U.S.
49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Arid and Semiarid
Arid and semiarid ecosystems are abundant in areas of the western U.S. where climate or
orography create annually or seasonally dry conditions. At the time of the 2008 ISA, a
large amount of information was available on how N deposition affected the growth and
physiology of plants and microorganisms in arid and semiarid ecosystems, and there was
coherent evidence that plant communities in these ecosystems could be altered by the
added N. Evidence for these effects was particularly strong in coastal sage scrub (CSS),
chaparral, and Mojave Desert ecosystems in southern California. Within the CSS
ecosystems, N deposition has been linked to increased mortality in native shrubs,
decreased abundance of arbuscular mycorrhizal fungi, higher cover of invasive annual
plants, and increased wildfire activity. Similar increases in invasive annual plant cover
and fire frequency have also been attributed to N deposition in areas of the Mojave
Desert downwind of urban centers in southern California. Research since 2008 has
further documented these effects, with consistent evidence that N deposition can affect
the physiology, growth, and community composition of plants and soil microorganisms
in arid and semiarid systems.
The effects of N deposition on physiological and biogeochemical processes in arid and
semiarid ecosystems are even more clearly dependent on moisture availability than in
grasslands (Appendix 6.2.6). In these ecosystems, inorganic N often accumulates in the
soil during dry periods, and growth and physiological responses to additional N are only
observed when and where sufficient moisture is available. Two additional important
effects of aridity include (1) higher soil base saturation and pH that buffer these systems
from the acidification effects of N deposition and (2) spatially patchy nutrient availability
that develops beneath isolated shrub canopies. One important effect of N deposition on
arid and semiarid ecosystems is to decrease the patchiness of nutrient availability, which
promotes the growth of invasive annual plants in the spaces between the isolated shrubs.
The growth of these annual plants creates a more continuous fuel bed for wildfires,
increasing the prevalence of fire, and shifting plant community composition toward more
fire-adapted plant species.
Since 2008, increases in aboveground plant biomass or plant cover have been observed in
the U.S. in the Mojave and Sonoran Deserts, and in southern California chaparral, and
internationally in China and Spain (Appendix 6.2.6). Given the linkage to fire, it is
notable that there have been multiple observations of increased annual plant growth in the
Mojave Desert in response to added N.
New research has also provided further evidence that N deposition alters plant
communities in arid and semiarid ecosystems, particularly in southern California, but also
in other locations (Appendix 6.3.6). Many of these studies documented changes in plant
50
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
community composition, with fewer observations of plant species loss or changes in plant
diversity. Overall, this body of research has provided consistent and coherent evidence
that N deposition is altering the growth, physiology, and community composition of
plants in arid and semiarid ecosystems. Relative to plants, there are fewer studies of
microbial communities (Table 6-12; Table 6-22). but these studies provided evidence that
N additions can alter the abundance, physiology, and community composition of soil
microorganisms in arid and semiarid ecosystems.
1.5.2.2 National-Scale Sensitivity and Critical Loads
At the time of the 2008 ISA, there had been little quantification of the extent and
distribution of N sensitivity in terrestrial ecosystems in the U.S. In the 2008 ISA, there
was no published U.S. national CL assessment. Since then, substantial work has been
done on quantifying N CLs for U.S. ecoregions. The most notable new work is the U.S.
Department of Agriculture—Forest Service (USDA—FS) Assessment of Nitrogen
Deposition Effects and Empirical Critical Loads (Pardo et al.. 201 la). That assessment
was organized by Level 1 ecoregions, and where data were available, CLs calculations
were made for individual ecosystem types (e.g., forests within the Mediterranean
California ecoregion) and life forms (i.e., lichens, mycorrhizal fungi). This ISA largely
follows that structure, reporting terrestrial N CLs for life forms (e.g., herbaceous plants)
within each ecoregion, which is a geographically defined area within a broader biome
(e.g., forests) based on distinct physical and biological features (e.g., Northwest Forested
Mountains, Eastern Temperate Forests, etc.).
In general, higher N CLs were often reported for regions with higher ambient N
deposition. One explanation for this pattern is that when ecosystems experience elevated
N deposition, the current condition already represents a change from the condition before
elevated N deposition (i.e., a pristine or near-pristine state). This pattern would explain
why the empirical CL is often above the ambient deposition even as that deposition
increases in the same ecosystem type across a region (Pardo etal.. 2011a).
Newer CL studies are presented in tandem with the CLs reported by Pardo etal. (2011a)
in Table 6-28 and Chapter 1, Figure 1-7. The majority of values for new CLs are within
the range of CLs identified by Pardo etal. (2011a). Notably, Simkin et al. (2016)
identified a new lower range of 7.9 kg N/ha/yr, and new lower CLs are denoted for alpine
ecosystems in the Northwest Forested Mountains ecoregion. There are also new CLs for
herbaceous species in two ecoregions that previously had no CL ITable 6-28. (Simkin et
al.. 2016)1.
June 2018
51
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
Temperate Sierras
*-<	Mediterranean California
"qS
>
oi
North American Deserts
O
'c
GJ
q	Great Plains
c
o
*3P Northwest Forested Mountains
3 O

4-8


>3 O
June 2018
52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
1.5.3
Biological Effects of Acidification
Since publication of the 2008 ISA, the overarching understanding of terrestrial
acidification has not appreciably changed. Recent research has confirmed and
strengthened this understanding and provided more quantitative information, especially
across regional-scale landscapes. A number of studies have evaluated the relationships
between soil chemistry indicators of acidification and ecosystem biological endpoints
(see Table 5-6). Soil chemistry indicators examined in recent literature include
exchangeable base cations (Be), soil pH, exchangeable acidity (H+ and Al), exchangeable
Bc:Al ratio, base saturation, and Al concentrations. The most common indicator used in
determining critical loads is the soil solution Bc:Al ratio. Appendix 5.2.1 discusses the
uncertainty considerations when using this indicator. Biological endpoints included in the
evaluations consisted of physiological and community responses of trees and other
vegetation, lichens, soil biota, and fauna.
1.5.3.1 Physiology and Growth
In the 2008 ISA, evidence was sufficient to infer a causal relationship between acidifying
deposition and changes in terrestrial biota; the evidence included changes in plant
physiology, plant growth, and terrestrial biodiversity. The physiological effects of
acidification on terrestrial ecosystems in the U.S. were well characterized at the time of
the 2008 ISA and included slower growth and increased mortality among sensitive plant
species. Consistent and coherent evidence from multiple species and studies in 2008
showed that the biological effects of acidification on terrestrial ecosystems were
generally attributable to physiological impairment caused by Al toxicity and decreased
ability of plant roots to take up base cations (Section 3.2.2.3 of the 2008 ISA). Much of
the new evidence for the negative effects of acidifying deposition comes from Ca
addition experiments, in which the addition of Ca has alleviated many of the negative
plant physiological and growth effects. Consistent with the findings of the 2008 ISA, the
body of evidence is sufficient to infer a causal relationship between acidifying N and
S deposition and the alteration of the physiology and growth of terrestrial organisms
and the productivity of terrestrial ecosystems.
In the 2008 ISA, acidifying deposition, in combination with other stressors, was found to
be a likely contributor to physiological effects that led to the decline of sugar maple (Acer
saccharum) trees occurring in portions of the eastern U.S. with base-poor soils. Studies
since the 2008 ISA support these findings (see Appendix 5.2.1.1). For example, recent
field studies have shown relationships between soil chemical indicator threshold values
and tree responses. Substantial declines in sugar maple regeneration have been found at
June 2018
53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
soil base saturation levels <20%, which is consistent with the range reported in the 2008
ISA.
In new studies, sugar maple showed positive growth and regeneration responses to
increasing exchangeable base cations, base saturation, and soil pH, and negative
relationships with increasing exchangeable Al. In other studies, the growth, regeneration,
and physiological responses of sugar maple to the soil conditions created by acidifying
deposition were reversed or ameliorated by Ca additions. Similarly, the 2008 ISA
reported that processes associated with soil acidification contributed to physiological
stress, high mortality rates, and decreasing growth trends of red spruce (Picea rubens)
trees. New evidence from Ca addition studies provides further support for these
mechanisms (see Appendix 5.2.1.2). Added Ca reversed or ameliorated many of the
physiological responses to acidification.
In the 2008 ISA, there was limited information on the effects of acidification on other
tree species. Since the 2008 ISA, research has observed varying physiological sensitivity
to soil acidification among eight eastern U.S. tree species. New studies since the 2008
ISA have also added new information about the effects of acidifying deposition on forest
understory vegetation, grasslands, lichens, and higher trophic level organisms (snails and
salamanders) that support the terrestrial acidification conclusions of the 2008 ISA.
1.5.3.2 Biodiversity
The 2008 ISA noted strong evidence that acidifying deposition could alter terrestrial
community composition and cause a loss of terrestrial biodiversity. The physiological and
growth effects of acidifying deposition are not uniform across species, resulting in altered
species composition and decreased biodiversity whereby sensitive species are replaced by
more tolerant species. For example, increasing soil base cation availability was tied to
greater sugar maple growth and seedling colonization, whereas American beech (Fagus
grandifolia) was relatively more dominant on soils with lower base cation availability
(see Appendix 5.2.1.3.1). Measurements of soil acid-base chemistry have been used as a
predictor of understory species composition, with 50 understory species associated with
high soil base cation status. In another set of studies, soil acid-base chemistry was
correlated with soil biodiversity and community composition. For example, addition of
Ca resulted in changes in soil bacterial community composition and bacterial community
structure that were correlated with soil exchangeable Ca, pH, and P (see Appendix 5.2.4).
Based on research included in the 2008 ISA and these new studies, the body of evidence
is sufficient to infer a causal relationship between acidifying N and S deposition and
June 2018
54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
the alteration of species richness, community composition, and biodiversity in
terrestrial ecosystems.
1.5.3.3 National-Scale Sensitivity and Critical Loads
The sensitivity of soils to acidifying deposition is discussed in detail in Appendix 4. In
general, surficial geology is the principal factor governing the sensitivity of terrestrial
ecosystems soil to acidification from S and N deposition. Other factors that contribute to
the sensitivity of soils to acidifying deposition include topography, soil chemistry, and
land use. Several widely accepted models are currently used in the U.S. to assess the
terrestrial effects of S and N deposition (Appendix 4.5). These models are typically used
to evaluate acidification effects on biota by assigning a value of a soil parameter that
relates to the onset of a harmful biological effect. Since the 2008 ISA, estimates of base
cation weathering (BCw), which are input to soil acidification models, have improved for
application in the U.S. and have yielded new critical loads. Forests of the Adirondack
Mountains of New York, Green Mountains of Vermont, White Mountains of New
Hampshire, the Allegheny Plateau of Pennsylvania, and mountain tops and ridges forest
ecosystems in the southern Appalachians are the regions that are most sensitive to
terrestrial acidification from atmospheric deposition (Appendix 3.2.4.2 of the 2008 ISA).
Models used to determine critical loads of acidifying deposition included SMB, STA,
MAGIC, ForSAFE-VEG, and empirical models. Several models and extrapolation
methods to estimate BCw rates were also investigated. The PROFILE model was
evaluated as a model to estimate soil BCw rates to support estimates of SMB critical
loads in the U.S. (see Appendix 4.5). In general, recently published models used soil
solution Bc:Al ranging from 1.0 to 10.0 as an indicator to estimate CLs in North
America.
Ecosystem sensitivities to ambient N and S deposition were also characterized by
developing CLs and exceedances (see Appendix 4.6: Chapter 1, Figure 1-8. and
Appendix 5.5). Calculated CLs for forest plots based on the soil solution Bc:Al of 10.0 in
the northeastern U.S. ranged from 11 to 6,540 eq/ha/yr (eq quantifies the supply of
available H+ ions, combining the acidifying effects of N and S deposition), and 15-98%
of these plot-level CLs were exceeded by N and S deposition. In this region, correlation
analyses showed that the growth of 17 tree species were negatively correlated with CL
exceedance. In Pennsylvania, CLs based on the soil solution Bc:Al of 10.0 for hardwood
forests ranged from 4 to 10,503 eq/ha/yr and were exceeded by estimated N and S
deposition in the Year 2002 in 53% of the plots. Several studies found that CL and
exceedance determinations could be influenced by BCw rates, soil chemical indicators, N
June 2018
55
DRAFT: Do Not Cite or Quote

-------
1
2
retention, tree species-specific base cation uptake, and/or bulk (i.e., wet) versus total
deposition 011 CL estimates.
Forest Ecosystems Critical Loads for Acidity
S + N eq/ha/yr
H 170- 1,000
¦I 1,001-2,000
2,001-4,000
| 4,001-6,000
| 6,001-8,800
States
No Data
eq = equivalent; ha = hectare; yr = year.
(A) McNulty et al. (2007) critical loads are mapped at 1-km2 grids (center map). For uncertainty, see Li and McNulty (2007).
(B.) Duarte et al. (2013) critical loads are mapped at 4-km2 grids; (C. and D.) Phelan et al. (2014) critical loads are mapped for each
sampling site (Pennsylvania). McDonnell et al. (2014b): Sullivan et al. (2011b): Sullivan et al. (2011a) critical loads are mapped as a
single point at the center point of the watershed (New York and North Carolina).
Source: http://nadp.sws.uiuc.edu/committees/clad/db/NCLDMapSummarv 2015.pdf.
Figure 1-8 Forest ecosystem critical loads for soil acidity related to base
cation soil indicators.
June 2018
56
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
1.6 Freshwater Ecosystem Nitrogen Enrichment and
Acidification
For freshwater systems, new evidence reinforces causal findings from the 2008 ISA
(Chapter 1, Table 1-1). New evidence also expands the scope of existing causal findings
to include additional biota affected by N enrichment and acidifying deposition, and
supports quantification of these effects with new critical loads (see Chapter 1.6.3.2).
Freshwater systems include lakes (lentic systems) and rivers and streams (lotic systems).
In freshwater ecosystems, N may cause N enrichment/eutrophication. Aquatic
eutrophication occurs as increased productivity of algae and aquatic plants, altered
nutrient ratios, and sometimes decreased oxygen levels. Deposition ofN, S, or N + S can
cause acidification, which affects watershed biogeochemical processes and surface water
chemistry. Freshwater N enrichment and acidification take place in sensitive ecosystems
across the U.S. at present levels of deposition and may occur simultaneously in some
water bodies.
New studies have added to the body of evidence in the 2008 ISA that N nutrient
enrichment and acidifying deposition alter freshwater biogeochemistry and subsequent
biological effects. There is new information on biogeochemical processes including
cycling ofN and S. BothN enrichment/eutrophication and acidification can impact
physiology, survival, and biodiversity of sensitive aquatic biota. The 2008 ISA and new
studies provide examples of lakes and streams that show signs of eutrophication,
especially increased algal growth and shifts in algal biodiversity, in response to N
addition. The current causal statement for nutrient enrichment effects ofN deposition
now includes altered algal growth and productivity as well as the endpoints of species
richness, community composition, and biodiversity reported in the 2008 ISA (Chapter 1,
Table 1-1). For biological effects of aquatic acidification, the current causal statement has
been expanded from the 2008 ISA to include the specific endpoints of physiological
impairment, alteration of species richness, community composition, and biodiversity
(Chapter 1, Table 1 -1). New studies also show that despite reductions in acidifying
deposition, many aquatic ecosystems across the U.S. are still experiencing changes in
ecological structure and functioning at multiple trophic levels. Although there is evidence
for chemical recovery in many previously acidified ecosystems, biological recovery has
been limited (Appendix 8.4).
A number of freshwater monitoring efforts have facilitated the analysis of long-term
trends in surface water chemistry and ecological response in areas affected by acidifying
(N + S) deposition (Appendix 7.1.3). Many of these studies have been conducted in the
U.S., especially in the Northeast and the Appalachian Mountains. Although many of
these monitoring programs were in existence at the time of the 2008 ISA and were
June 2018
57
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
considered in that analysis, more recent publications reflect the longer period of
monitoring record and strengthen previous conclusions. Surface water chemistry data
from long-term monitoring by federal, state, and local agencies as well as university
research groups and nonprofits has been combined into several publicly available
metadatabases (Appendix 7.1.3.2) enabling further regional trend analysis. Since the
early 2000s, U.S. EPA, together with the states, tribes, and other entities and individuals,
have collaborated on a series of statistically representative surveys (National Aquatic
Resource Surveys [NARS]) of the nation's waters, including surveys of lakes (U.S. EPA.
2016h. 2009b). streams (U.S. EPA. 2016i). wetlands (U.S. EPA. 2016j). and coastal
waters (U.S. EPA. 2016g). These periodic surveys, which are based on standard sampling
and analysis protocols and consistent quality assurance, include chemical and biological
indicators of nutrient enrichment and acidification (Appendix 7.1.3).
1.6.1 Freshwater Biogeochemistry
In the 2008 ISA, evidence was sufficient to infer a causal relationship between N and S
deposition and the alteration of biogeochemical cycling of N and C in freshwater
ecosystems, and between acidifying deposition and changes in biogeochemistry of fresh
waters. As documented in the 2008 ISA and by newer studies, biogeochemical processes
and surface water chemistry are influenced by characteristics of the catchment and the
receiving waters. A number of studies since 2008 have focused on improving
understanding of aquatic acidification and eutrophication processes mediated by N. Many
of these studies have focused on pathways of pollutant and other constituent movement
within ecosystems, including monitoring studies of various kinds. Chemical indicators of
N deposition identified by the 2008 ISA were NOs and DIN concentrations in surface
waters. Increased N deposition to freshwater systems via runoff or direct atmospheric
deposition, especially to N limited and N and phosphorus (P) colimited systems, can alter
N cycling (Appendix 7) and stimulate primary production (Appendix 9). Data from
long-term monitoring, experimental manipulations, and modeling studies provide
consistent and coherent evidence for biogeochemical changes associated with acidifying
N and S deposition. The strongest evidence for a causal relationship between acidifying
deposition and aquatic biogeochemistry comes from studies of changes in surface water
chemistry. Surface water chemistry indicators of acidic conditions and acidification
effects include concentrations of SO42 . NO3 . inorganic aluminum (Al), calcium (Ca),
sum and surplus of base cations, acid-neutralizing capacity (ANC), and surface water pH.
New information on biogeochemical cycling of N and S, acidifying deposition effects on
biogeochemical processes and changes to chemical indicators of surface water chemistry
associated with acidification and N nutrient enrichment is consistent with the conclusions
June 2018
58
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
of the 2008 ISA, and the body of evidence is sufficient to infer a causal relationship
between N and S deposition and the alteration of freshwater biogeochemistry.
1.6.1.1 Freshwater Processes and Indicators
Key processes and geochemical indicators of freshwater acidification and N enrichment
(Chapter 1, Table 1-3) link to biological effects (Appendix 8 and Appendix 9). Surface
water chemistry integrates the sum of soil and water processes that occur upstream within
a watershed. Several key biogeochemical processes cause or contribute to surface water
eutrophication and acidification, and these processes have been the focus of substantial
research over the last three decades. Since the 2008 ISA, experimental studies, isotopic
analyses, and monitoring and observational studies have further investigated the cycling
of S, N, C, and base cations; these studies substantiate and further quantify earlier
findings.
Nitrogen is deposited as NOs . NH4 . NH3, and/or organic N. Inorganic N is leached from
terrestrial ecosystems mainly as NO3 . In freshwater ecosystems, deposited NH4 is taken
up by biota or nitrified to NO3 . Elevated NO3 concentrations in lakes and streams are a
biogeochemical indicator that a freshwater system is receiving excess N which will cause
acidification or eutrophication. Qualitatively, northeastern U.S. spatial patterns in surface
waters NO3 concentrations suggest an influence by atmospheric N deposition. However,
considerable variation in the relationship between stream chemistry and deposition was
associated with land use and watershed attributes. It was well known at the time of the
2008 ISA that key processes such as nitrification and denitrification are quantitatively
important portions of the N cycle and that they can be influenced by atmospheric inputs.
More recent research has further substantiated these earlier findings and provided
additional quantitative context (Appendix 7.1.2.3). Some new research suggests that
denitrification may, in some situations, produce more N2O in relationship to surface
water NO3 concentration than was previously recognized.
Sulfur is deposited mainly as SO42 , which is a mobile anion in many acid-sensitive
watersheds (Appendix 4). Deposition is not the only source of SO42 to drainage waters.
Geologic sources of S, including iron sulfide minerals, can also contribute SO42 to
surface waters. The 2008 ISA found that S deposition alters soil and drainage water
chemistry through sustained leaching of SO42 , associated changes in soil chemistry, and
accumulation of S in the soil through adsorption and biological assimilation. Declines in
lake S042 concentrations have been observed in locations where S deposition has
decreased significantly such as in the Adirondack Mountains (Appendix 7.1.5.1). In
addition, internal watershed sources of S, which were earlier believed to be relatively
June 2018
59
DRAFT: Do Not Cite or Quote

-------
1	minor in the northeastern U.S., have and will likely continue to become proportionately
2	more important as S deposition continues to decline. Reductions in SOx deposition have
3	not consistently resulted in increases of ANC in surface water.
Table 1-3 Summary of key aquatic geochemical processes and indicators
associated with eutrophication and acidification.
Endpoint
N Driven
Nutrient
Enrichment
Acidification
The Effect of Deposition
Process
NOs"
leaching into
water bodies
X
X
Leaching from terrestrial ecosystems is an important source of NO3 in
freshwater ecosystems. See NO3" leachinq in Chapter 1, Table 1-2.
so42~
leaching into
water bodies

X
Leaching from terrestrial ecosystems is an important source of SC>42~
in freshwater ecosystems. See SO42" leachinq in Chapter 1, Table 1-2.
Nitrification
X
X
Nitrification is an acidifying process, releasing 2 mol hydrogen ion (H+)
per mol NhV converted to NO3". As the N cycle becomes enriched
through cumulative N addition, net nitrification rates often increase,
and N03"concentrations increase.
Denitrifi cation
X

Denitrification is the microbial process that transforms NO3" by
anaerobically reducing it to NO2", NO, N2O, and N2.
DOC
leaching into
water bodies
X
X
DOC contributes to acidity of freshwater ecosystems. See DOC
leachinq in Chapter 1. Table 1-2.
Indicator
Surface
water [NO3"]
X
X
Increased N deposition (to surface waters or to terrestrial watershed,
see Chapter 1. Table 1-2) increases the water NO3" concentration.
High concentrations of NO3" in lakes and streams, indicative of
terrestrial ecosystem N saturation, have been found at a variety of
locations throuqhout the U.S. (U.S. EPA. 2006c: Stoddard. 1994).
Comparison of preindustrial to modern estimates suggested elevated
concentrations in water bodies as a result of N deposition (Fenn et al..
2011b).
Surface
water DIN
X

Increased N deposition increases DIN in most freshwater aquatic
environments, largely as NO3"
Surface
water N:P
ratios
X

Increased N deposition can alter the ratio of N to P in freshwater
systems. Freshwater biota have different nutrient requirements and
changes in nutrient ratios may alter species richness, community
structure, and biodiversity, especially primary producers.
June 2018
60
DRAFT: Do Not Cite or Quote

-------
Table 1-3 (Continued): Summary of key aquatic geochemical processes and
indicators associated with eutrophication and acidification.
N Driven
Nutrient
Endpoint Enrichment Acidification	The Effect of Deposition
Surface	X	Increased S deposition (to surface waters or to terrestrial watershed,
water [SO42"]	see Chapter 1, Table 1-2) increases the water SO42" concentration.
Comparison of preindustrial to modern estimates suggested elevated
concentrations in water bodies as a result of S deposition.
Several studies in the eastern U.S. suggested that base cation
concentrations in surface waters increased during the initial phases of
acidification into the 1970s. This trend reversed, and base cations
have decreased in response to decreasing SO42" and NO3"
concentrations. Many base cations (especially Ca2+) are important
nutrients for aquatic biota.
Increased N and S deposition decrease ANC. Surface water ANC
correlates with other biologically influential chemical metrics, including
pH, inorganic Al concentration, Ca concentration, and organic acidity.
ANC <50-100 peq/L typically poses a risk for species survival, species
richness, and biodiversity.
Surface water pH is a common alternative to ANC as an indicator of
acidification, but ANC is a better indicator at pH >6.0, and is less
sensitive to dissolved CO2. N and S deposition are associated with
decreasing pH in surface waters. Increasing pH trends in surface
waters in the northeastern U.S. were common through the 1990s up to
2004, but the rates of change have been small (Driscoll et al.. 2007a:
Driscoll et al.. 2001a; Driscoll et al.. 2001b).
X	Acidifying N and S deposition increase mobilization of inorganic Al
from terrestrial ecosystems into surface water, increasing surface
water concentrations. Inorganic Al in surface waters is (1) widely toxic
and (2) leaches from terrestrial ecosystems only in response to acidic
conditions. Earlier studies demonstrated reduced growth and survival
of various species offish at inorganic Al concentrations between
approximately 2 and 7.5 pmol/L. Most recently, 20% mortality of
young-of-the year brook trout was documented in situ during a 30-day
period with a median inorganic Al concentration of 2 pmol/L.
Al = aluminum; ANC = acid-neutralizing capacity; Ca = calcium; C02 = carbon dioxide; DIN = dissolved inorganic nitrogen;
DOC = dissolved organic carbon; H+ = hydrogen ion; ha = hectare; kg = kilogram; L = liter; |jeq = microequivalents;
|jmol = micromole; N = nitrogen; N2 = molecular (atmospheric) nitrogen; N20 = nitrous oxide; NE = northeast; NH4+ = ammonium;
NO = nitric oxide; N02 = nitrogen oxide; N03" = nitrate; P = phosphorus; S = sulfur; S042" = sulfate; U.S. = United States;
USFS = U.S. Forest Service; yr = year.
1.6.1.1.1	Acidification
2	The acidifying effects of N and S deposition in U.S. waters have been well characterized
3	for several decades. Traditionally acidification involves both chronic and episodic
4	processes. Driscoll et al. (200lb) characterized chronically acidic lakes and streams by
5	ANC of <0 |icq/L throughout the year, while episodic acidification occurs when ANC
June 2018	61	DRAFT: Do Not Cite or Quote
Surface	X
water (base
cation)
Surface	X
water ANC
Surface	X
water pH
Surface
water
Inorganic Al

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
falls below 0 (j,eq/L only for hours to weeks. Chronic acidification refers to average
conditions and is often measured as summer and fall chemistry for lakes and as spring
baseflow chemistry for streams. Chronic conditions are no longer prevalent in regions of
the U.S. affected by acidic deposition (cf., Fakhraei et al.. 2016; Fakhraei et al.. 2014).
Episodic acidification is associated with precipitation or snowmelt events when high
volumes of drainage water enter watersheds. Episodes generally cause changes in at least
two of the following chemical parameters: ANC, pH, base cations, SO42 concentration,
N03 concentration, inorganic Al concentration, organic acid anions, or DOC. New
studies show that both N and S contributed to episodic acidification over a 20-year period
at Bear Brook, ME (see Appendix 7.1.5.1.2). It is known that the biota in many
streams/lakes are impacted when the ANC is consistently below 50 ueq/L. For this
reason, there has been a shift away from focusing on chronic versus episodic condition.
For example, the U.S. EPA National Lakes Assessment used an ANC threshold of
>50 ueq/L as indicative of nonacidified water bodies (U.S. EPA. 200%).
The most widely used measure of surface-water acidification is ANC. As reported in the
2008 ISA and newer studies, ANC is the primary chemical indicator of historic
acidification and for predicting the recovery expected from decreasing atmospheric
deposition. ANC correlates with the surface water constituents (pH, Ca2+, and inorganic
Al concentration) that contribute to or ameliorate acidity effects in biota. As reported in
the 2008 ISA, lake and stream ANC values decreased throughout much of the 20th
century in a large number of acid-sensitive lakes and streams throughout the eastern U.S.
This effect has been well documented in monitoring programs, paleolimnological studies,
and model simulations (Appendix 7.1.5.1). Biological indicators of acidification, such as
decreased fish species richness, are presented in Appendix 8.3.
Surface water pH is another indicator of acidification. It also correlates with surface
water chemical constituents with biotic effects (inorganic Al, Ca2+, and organic acids).
The 2008 ISA included the scientific consensus that low pH can have direct toxic effects
on aquatic species (U.S. EPA. 2008a; Driscoll et al.. 2001b). A pH value of 6.0 is the
level below which biota are at increased risk from acidification (Appendix 8.3). In the
2008 ISA, increasing trends in pH (decreasing acidification) were common in surface
waters in the northeastern U.S. through the 1990s and up to 2004 and have continued in
more recent times at many locations (Appendix 7.1.2.5). Rates of change have generally
been relatively small.
As stated in the 2008 ISA, the concentration of dissolved inorganic monomeric Al in
surface waters is an especially useful indicator of acidifying deposition because (1) it is
toxic to many aquatic species and (2) it leaches from soils only under acidic conditions
including acidifying deposition, acid mine drainage, or from rare geologic deposits.
62
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Inorganic A1 has well-documented effects on aquatic biota at specific thresholds
(Appendix 8.3) and is often the greatest threat to aquatic biota below pH 5.5. The 2008
ISA noted that concentrations of inorganic A1 decreased slightly in some surface waters
in the northeastern U.S. in response to decreased levels of acidifying deposition
suggestive of chemical recovery of surface waters (U.S. EPA. 2008a). and this trend has
generally continued (Appendix 7.1.5. see discussion on recovery Chapter 1.11).
Assessments of acidifying deposition effects dating from the 1980s and reported in the
2008 ISA showed SOr to be the primary acidifying ion in most acid-sensitive waters in
the U.S. The 2008 ISA, presented temporal data that showed a trend of increasing
concentrations of SOr in surface waters before the period of peak S emissions in the
early 1970s. After the peak, SO42 surface water concentrations decreased in a
widespread trend. The rate of recovery varied by ecosystem and new studies indicate that
as atmospheric S deposition has declined, soils with large stores of historically deposited
S (e.g., the Blue Ridge Mountain region) have begun releasing this adsorbed S to
drainage water, (Appendix 4) preventing or slowing aquatic recovery.
As stated in the 2008 ISA, the quantitatively most important component of the overall
surface water acidification and chemical recovery responses has been change in base
cation supply. Decreases in base cation concentrations in surface waters in the eastern
U.S. have been ubiquitous over the past two to three decades and closely tied to trends in
S042 concentrations in surface waters. Change in base cation supply with surface water
acidification was highlighted in the assessment of Charles and Christie (1991) and in the
2008 ISA. Base cations are added to watershed soils by weathering of minerals and
atmospheric deposition, and are removed by uptake into growing vegetation or by
leaching. Acidic deposition increased leaching of base cations, as SO42 anions in soil
solution carried along base cations to maintain the charge balance. In watersheds that
received high levels of historical acidic deposition, current exchangeable concentrations
of Ca2+ and other base cations are substantially reduced from likely preindustrial levels,
having been depleted by many years of acidic deposition. This base cation depletion in
watersheds constrains ANC and pH recovery of surface waters, as described in the 2008
ISA. New studies of base cations have further corroborated these earlier findings and
included experiments, modeling, and gradient studies.
Changes in DOC concentration or properties can impact the acid-base chemistry of
surface waters and perhaps the composition of aquatic biota. It has been recognized that
surface water DOC concentrations had decreased to some extent as a result of
acidification, and that DOC would likely increase with recovery. However, the strength
of this response and the magnitude of DOC changes have exceeded scientific predictions.
Recent research on this topic has been diverse and has included experiments, observation,
63
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
isotope studies, and synthesis and integration work. Overall, these studies illustrate large
increases in DOC with acidification recovery in some aquatic systems. Increases in DOC
constrain the extent of ANC and pH recovery, but decrease the toxicity of dissolved A1
by converting some of it from inorganic to organic forms (Lawrence et al.. 2013).
However, DOC is not an indicator of recovery everywhere; some recovering sites have
not shown increasing trends in DOC.
Nitrogen Enrichment/Eutrophication
In aquatic systems, N is a nutrient that stimulates growth of primary producers (algae
and/or aquatic plants). Atmospheric deposition of N to freshwater systems can increase
the absolute supply of nutrients and alter N and P ratios. The freshwater ecosystems in
the U.S. most likely to be sensitive to nutrient enrichment from N deposition are
headwater streams, lower order streams, and alpine lakes, which have very low nutrients
and productivity and are far from local pollution sources IYU.S. EPA. 2008a).
Appendix 9.1.1.41. These nutrient shifts alter stoichiometric composition of water
chemistry, thereby shifting the nutrient status of lakes. Even small inputs of N in low
nutrient water bodies can affect biogeochemical processing of N and increase the
productivity of photosynthesizing organisms, resulting in a larger pool of fixed carbon
(C). Nutrient enrichment leads to changes in aquatic assemblages and biodiversity in
freshwater (Appendix 9) and coastal regions (Appendix 10).
Indicators of altered N cycling include changes in the concentrations of NOs in surface
waters. The concentration of NO3 in drainage water provides an index of the balance
between removal and addition of N to terrestrial ecosystems. Studies of several types
have been conducted in recent years to elucidate these processes and include
experimental studies, isotopic analyses, and monitoring and observational studies. Both
water column and sediment N transformations have been further characterized
(Appendix 7.1.2.3). New research since the 2008 ISA suggests that denitrification may,
in some situations, play a larger role than was previously recognized in removing
oxidized N from the watershed.
As reported in the 2008 ISA and in newer studies, atmospheric N has been positively
correlated to total N in lakes along gradients of atmospheric deposition and N deposition
in some high-deposition lakes has changed the nutrient status of these lakes from a more
or less balanced (mainly N deficient) state to more consistently P limited conditions
(Appendix 9.2.4). Since the 2008 ISA, several studies have reported increases in P
deposition to water bodies in the U.S., possibly effecting shifts from N to P limitation or
colimitation, as well as P deposition prolonging N limitation (Appendix 9.1.1.2). In
higher order streams, N deposition typically mixes with N derived from other
64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
nonatmospheric sources, including urban/suburban point and nonpoint sources, industrial
sources, and agricultural sources, with atmospheric sources typically being most
pronounced during high-flow conditions (Table 7-2).
1.6.1.2 Models
Models used to assess the effects of N and S deposition on U.S. ecosystems were
reviewed in the 2008 ISA (Annex A). Several of the models used for terrestrial
ecosystems (see Chapter 1.5.3.3) such as MAGIC and PnET/BGC are also applicable to
aquatic systems. Both of these models have been widely applied, mainly to relatively
small, upland watersheds. Three other models, Spatially Referenced Regressions on
Watershed Attributes (SPARROW), Watershed Assessment Tool for Evaluating
Reduction Scenarios for Nitrogen (WATERS-N), and Surface Water Assessment Tool
(SWAT) have been used to evaluate N loading to mixed-use watersheds in larger river
systems. A model that has been applied to the analysis of nutrient enrichment in aquatic
systems is AQUATOX, which simulates nutrient dynamics and effects on aquatic biota.
Few new freshwater acidification or eutrophication models have been developed and
published since 2008. A new national water quality modeling system (Hydrologic and
Water Quality System, HAWQS) is under development by Texas A&M University and
the USDA for the U.S. EPA's Office of Water (https: //epahawq s .tamu. edu/). The model
is intended to assist resource managers and policy makers in evaluating the effectiveness
of water pollution control efforts. Freshwater eutrophication and acidification models are
described in greater detail in Appendix 7.1.4.2.
1.6.1.3 National-Scale Sensitivity
Sensitivity of lakes, streams, and rivers to biogeochemical changes associated with N and
S deposition varies across the U.S. The biogeochemical sensitivity to acidifying
deposition will be discussed together with biological sensitivity in Chapter 1.6.2.2.
Sensitivity to N enrichment will be discussed with biological sensitivity in
Chapter 1.6.3.2.
1.6.2 Biological Effects of Freshwater Nitrogen Enrichment
In the 2008 ISA, evidence was sufficient to infer a causal relationship between N
deposition and the alteration of species richness, community composition, and
biodiversity in freshwater ecosystems. The freshwater systems most affected by nutrient
June 2018
65
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
enrichment due to atmospheric deposition of N were remote oligotrophic high-elevation
lakes with low N retention capacity. In these ecosystems, N changes the biota, especially
by increasing algal growth and shifting algal communities. Freshwater organism
responses to N enrichment can be assessed through biological indicators, including
chlorophyll a, phytoplankton and periphyton (algae attached to a substrate) biomass,
diatoms, and trophic status. The current causal statement has been expanded to include
effects on algal growth and productivity (Chapter 1, Table 1-1). New evidence since 2008
of N enrichment includes paleolimnology, phytoplankton community dynamics,
macroinvertebrate response, and indices of biodiversity. This new evidence is consistent
with the conclusions and strengthens the evidence base of the 2008 ISA, and together, the
body of evidence is sufficient to infer a causal relationship between N deposition and
changes in biota including altered growth and productivity, species richness,
community composition, and biodiversity due to N enrichment in freshwater
ecosystems.
1.6.2.1 Physiology and Biodiversity Effects
Inputs of N to freshwater systems stimulate algal growth, which leads to a cascade of
effects on algal community composition and biodiversity. Algal species have differential
responses to N loading and shifts in nutrient ratios, so dominant species may change in
response to N enrichment. As reported in the 2008 ISA and in newer studies, shifts in
nutrient limitation from N limitation to colimitation by N and P, or to P limitation, have
been observed in some alpine lakes. New biodiversity studies are summarized in
Table 9-3. Since the 2008 ISA, several meta-analyses have reported an increase in P
deposition to water bodies, highlighting the need to account for how sustained P
deposition can modify the effects of anthropogenically emitted N deposition on
productivity (Appendix 9.1.1.4). P addition delayed the shift to P limitation (prolonged N
limitation) for phytoplankton.
1.6.2.1.1	Primary Producers
The body of evidence for biological effects ofN enrichment in remote freshwater
systems (where atmospheric deposition is the predominant source of N) is greatest for
phytoplankton, the base of the freshwater food web. The majority of studies focused on
phytoplankton, although several new studies indicate that both benthic and pelagic
primary producers respond to N inputs, and at least some studies have shown that
periphyton outcompeted phytoplankton for limiting nutrients (Appendix 9.3.3). The 2008
ISA and new studies include lake surveys, fertilization experiments, and nutrient
June 2018
66
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
bioassays that show a relationship between increased N concentrations and increased
pelagic and benthic algal productivity (measured by chlorophyll a concentration). An
increase in lake phytoplankton biomass with increasing N deposition was reported in the
Snowy Range in Wyoming and in Europe. New studies in the Colorado Rocky
Mountains, where atmospheric deposition ranged from 2 to 7 kg N/ha/yr, found
correlations between higher chlorophyll a and higher rates of deposition
(Appendix 9.2.1).
The 2008 ISA and newer studies (Table 9-1 and Appendix 9.3.2) show a general shift in
algal dominance from chrysophytes that dominate low N lakes to cyanophytes and
chlorophytes in higher N lakes. Two nitrophilous species of diatom, Asterionella formosa
and Fragilaria crotonensis, serve as indicators of N enrichment in lakes; however,
increased relative abundance of A. formosa has also been attributed to lake warming in
some regions where N deposition is decreasing (Appendix 9.2.3). New studies show that
glacial meltwater has higher NOs relative to snow meltwater with different influences on
algal community composition in some regions of the U.S. (Appendix 9.3.2). In a
comparison of lakes in the Rockies with different meltwater sources, fossil diatom
richness in snowpack-fed lakes was at least double the richness of lakes with both glacial
and snow meltwater inputs; however, alterations in phytoplankton community structure
were not observed in lakes in the Northern Cascade Mountains, WA. Some shifts in algal
biodiversity observed in high-elevation waters are attributed to climate change or nutrient
effects and climate as costressors (Appendix 13).
Since the 2008 ISA further studies have shown that both trophic interactions and DOC
modify ecosystem response to N loading. DOC affects acidity and N cycling and is
increasing in some U.S. surface waters (Appendix 7.1.2.9). In a whole lake N fertilization
study (Appendix 9323), observed changes in community composition of phytoplankton
were related to DOC rather than N addition to small N limited boreal lakes.
The role of N in freshwater HAB formation has been further researched since the
2008 ISA. Additional evidence continues to show that availability and form of N
influences algal bloom composition and toxicity, and inputs of inorganic N selectively
favor some HAB species, including those that produce microcystin. Microcystin is
prevalent in U.S. waters as reported in recent regional and national surveys. Although the
risk of HAB formation is low in high-elevation oligotrophic water bodies where N
deposition is the dominant source of N, transport of atmospheric inputs can exacerbate
eutrophic conditions in downstream water bodies. Increased understanding of the role of
N as a limiting nutrient in many freshwater systems has led to recommendations to
consider both N and P in nutrient-reduction strategies.
67
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
Few studies in the U.S. have considered the effects of atmospheric deposition on aquatic
macrophytes, although declines in macrophyte occurrence were noted in a new survey of
Lake Tahoe that compared the lake's biota with that from a survey conducted in the
1960s (Caires et al.. 2013). Atmospheric N contributions are a substantial portion
(approximately 57%) of the total N loading to Lake Tahoe.
1.6.2.1.2	Zooplankton
Compared to changes in primary producers, biological responses to N deposition at
higher trophic levels are not well characterized, but atmospheric N can alter food web
interactions (see Appendix 9.3.4). A few studies in the 2008 ISA and newer studies
showed zooplankton responses to N related shifts in phytoplankton biomass potentially
altering food web interactions.
1.6.2.1.3	Macroinvertebrates
Only a limited number of studies published since the 2008 ISA have linked atmospheric
N deposition to taxonomic shifts and declines in invertebrates (Appendix 9.3.5). These
studies do not attribute shifts in the abundance of higher invertebrates to N deposition
alone, as there are interactions with climate and invasive species. New studies provide
additional evidence that trophic interactions may moderate algal growth following
nutrient loading. In Lake Tahoe, which receives 57% of N inputs from atmospheric
sources, endemic invertebrate taxa have declined 80 to 100% since the 1960s due to
nutrient inputs and invasive species.
1.6.2.2 National-Scale Sensitivity and Critical Loads
New data have not appreciably changed the identification of sensitive lakes and streams
in the U.S. Nutrient enrichment effects from N most likely occur in undisturbed,
low-nutrient headwater, and lower order streams and lakes at higher elevations in the
western U.S. (Appendix 9.1). including the Snowy Range in Wyoming, the Sierra
Nevada, and the Colorado Front Range. A portion of these lakes and streams where
effects are observed are in Class I wilderness areas. The responses of high-elevation lakes
vary with catchment characteristics (Appendix 9.1) and N deposition estimates at these
high elevation sites are associated with considerable uncertainty, and there is greater
uncertainty for estimates of dry deposition (Appendix 2). In these systems, even low
June 2018
68
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
inputs of atmospheric N can shift N limitation to colimitation by N and P, or to P
limitation (Appendix 9.2.4). altering algal species composition and productivity.
In the 2008 ISA, diatom assemblage shifts were reported at N deposition rates as low as
1.5 kg/N/yr. Additionally, a hindcasting exercise in remote alpine Rocky Mountain
National Park lakes associated algal changes between 1850 and 1964 with an increase in
wet N deposition of 1.5 kg N/ha/yr. Since the 2008 ISA, empirical and modeled CLs for
the U.S. have been estimated based on surface water NOs concentration, diatom
community shifts, and phytoplankton biomass nutrient limitation shifts indicative of a
shift from N limitation to P limitation. A critical load ranging from 3.5 to 6.0 kg N/ha/yr
was identified for high-elevation lakes in the eastern U.S. based on the nutrient
enrichment inflection point [where NOs concentrations increase in response to
increasing N deposition (Baron et al.. 201 lb)I. Another critical load of 8.0 kg N/ha/yr for
eastern lakes based on the value of N deposition at which significant increases in surface
water NO3 concentrations occur was estimated by (Pardo et al.. 2011c). In both Grand
Teton and Yellowstone National Parks critical loads for total N deposition ranged from
<1.5 + 1.0 kg N/ha/yr to >4.0 + 1.0 kg N/ha/yr (Nanus et al.. 2017). Exceedance
estimates were as high as 48% of the Greater Yellowstone area study region, depending
on the threshold value of NO, concentration in lakewater selected as indicative of
biological harm. An empirical CL of 4.1 kg N/ha/yr above which phytoplankton biomass
P limitation is more likely than N limitation was identified by Williams et al. (2017b) for
the western U.S. Modeled critical loads ranged from 2.8 to 5.2 kg/N/ha/yr, and a
performance analysis indicated that a CI of 2.0 kg N/ha/yr would likely reduce the
occurrence of false negatives to near zero.
1.6.3 Biological Effects of Freshwater Acidification
The 2008 ISA found evidence sufficient to infer a causal relationship between acidifying
deposition and changes in aquatic biota, including strong evidence that acidified aquatic
habitats had lower species richness of fishes, macroinvertebrates, and phytoplankton. The
effects of acidifying deposition on aquatic ecosystems also include physiological
impairment or mortality of sensitive species and shifts in biodiversity of both flora and
fauna. Organisms at all trophic levels are affected by acidification, with clear linkages to
chemical indicators for effects on algae, benthic invertebrates, and fish (Table 8-9).
Biological effects are primarily attributable to low pH and high inorganic Al
concentration. ANC integrates chemical components of acidification (Chapter 1,
Table 1-2). and surface water acidification models project ANC rather than pH and
inorganic Al concentrations. However, ANC does not directly alter the health of biota.
June 2018
69
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Effects of acidification on fish species are especially well characterized and many species
are harmed. Both in situ and lifestage experiments in fish support thresholds of chemical
indicators for biological effects. Most of these effects were documented in a rigorous
review of acidification effects on aquatic biota that was included in the 2008 ISA.
Overall, the updated research synthesized in this ISA reflects incremental improvements
in scientific knowledge of aquatic biological effects and indicators of acidification as
compared with knowledge summarized in the 2008 ISA. The fundamental understanding
of mechanisms has not changed, and the causal relationships between acidifying
deposition and biological effects on aquatic ecosystems are now, and were in 2008, well
supported. New studies also show that despite reductions in acidifying deposition,
alterations in aquatic biodiversity and ecosystem functioning caused by acidification
persist. Although there is evidence for chemical recovery in many ecosystems, biological
recovery has been limited (Chapter 1.6.2.2). New research is consistent with the causal
determination in the 2008 ISA and has strengthened the evidence base for these effects.
The current causal statement has been expanded to include specific endpoints of
physiological impairment, as well as effects at higher levels of biological organization
(Chapter 1, Table 1-1). The body of evidence is sufficient to infer a causal relationship
between acidifying deposition and changes in biota including physiological
impairment and alteration of species richness, community composition, and
biodiversity in freshwater ecosystems.
1.6.3.1 Physiology and Biodiversity Effects
The deterioration in water quality caused by acidification affects the physiology,
survivorship, and biodiversity of many species from several taxonomic groups and at
multiple trophic levels. As stated in the 2008 ISA, biological effects are primarily
attributable to low pH (or ANC) and high inorganic A1 concentrations under chronic or
episodic acidic conditions. During acidification episodes, water chemistry may exceed
the acid tolerance of resident aquatic biota, with effects that include fish mortalities,
changes in species composition, and declines in species richness across multiple taxa.
Studies reviewed in the 2008 ISA showed that the earlier aquatic lifestages were
particularly sensitive to acidification. New effects thresholds have been identified for
aquatic organisms consistent with observations in the 2008 ISA (Table 8-10). New
evidence is congruent with findings in the 2008 ISA that high levels of acidification (to
pH values below 5 and ANC lower than the range of 50 to 100 (j,eq/L) eliminate sensitive
species from freshwater streams. This information is reviewed below.
June 2018
70
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Primary Producers
Phytoplankton are primary producers at the base of the aquatic food web. These
photosynthetic organisms vary in tolerance of acidic conditions and include diatoms,
cyanobacteria, dinoflagellates, and other algal groups. The 2008 ISA reported reduced
species richness of freshwater plankton in response to acidification-related decreases in
pH and increases in inorganic Al. Effects were most prevalent when pH decreased to the
5 to 6 range. Effects on productivity are uncertain. Since the 2008 ISA, several
paleolimnological and field studies have further linked phytoplankton community shifts
to chemical indicators of acidification (Appendix 8.3). For example, Lacoul et al. (2011)
reviewed information on the effects of acidification and observed that the largest declines
in phytoplankton species richness occur over a pH range of 4.7 to 5.6 in Atlantic Canada.
Zooplankton
Zooplankton, the animal forms of plankton, comprise many groups of freshwater
unicellular and multicellular organisms including protozoans, rotifers, cladocerans, and
copepods. Zooplankton feed on phytoplankton or other zooplankton. Decreases in ANC
and pH and increases in inorganic Al concentration have been shown to contribute to the
loss of zooplankton species or abundance in lakes. In the 2008 ISA, thresholds for
zooplankton community alteration were between pH 5 and 6. In the Adirondacks, a
decrease in pH from 6 to 5 decreased zooplankton richness in lakes, and at ANC <0,
zooplankton richness was only 45% of the richness in unacidified lakes. Newer studies
support effects in a similar pH range (see Appendix 8.3.1.2). Zooplankton have also been
used as indicators of biological recovery (Appendix 8.4.2).
Benthic Invertebrates
Acidification has strong impacts on aquatic invertebrates because H+ and Al are directly
toxic to sediment-associated invertebrates like bivalves, worms, gastropods, and insect
larvae. In the 2008 ISA and in new studies in Appendix 8.3.3. decreases in ANC and pH
and increases in inorganic Al concentration contribute to declines in abundance or
extirpation of benthic invertebrate species in streams. Acidification to pH values below 5
eliminates mayflies (Ephemeroptera), a taxa indicative of stream water quality, along
with other aquatic organisms. Since the 2008 ISA, a survey of benthic macroinvertebrates
by Baldigo et al. (2009) in the Adirondack Mountains indicated that macroinvertebrate
communities were intact at a pH above 6.4, with moderate acidification effects at pH 5.1
to 5.7, and severe acidification effects at a pH less than 5.1. Similarly, thresholds of pH
71
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
5.2 to 6.1 were identified for sensitive invertebrates from Atlantic Canada
(Appendix 8.3.3V
1.6.3.1.4 Fish
The effects of low pH and ANC and of high inorganic A1 concentrations have been well
characterized in fish for many decades (Appendix 8.3.6). The 2008 ISA reported that
acidification impairs gill function and can cause respiratory and circulatory failure in fish.
Sensitivity to pH and inorganic A1 varies among fish species, and among lifestages within
species, with early lifestages more sensitive to acidification. The most commonly studied
species were brown trout (Salmo trutta), brook trout (Salvelinus fontinalis), and Atlantic
salmon (Salmo salar). Studies published since the 2008 ISA, especially in Atlantic
salmon, add to the existing information on sublethal effects confirming variation in
sensitivity among lifestages (Appendix 8.3.6.IV Since 2008, new studies include
acidification effects upon migratory activities and upon behavior. New studies on fish
show behavioral effects at pH <6.6 (Appendix 8.3.6.5).
As summarized in Baker etal. (1990a) and the 2008 ISA, fish populations in acidified
streams and lakes of Europe and North America have declined, and some have been
eliminated as a result of atmospheric deposition of N and S and the resulting changes in
pH, ANC, and inorganic A1 concentrations in surface waters. There is often a positive
relationship between pH and the number of fish species, particularly between pH 5.0 and
6.5. Additional pH thresholds published since the 2008 ISA (Table 8-2) support this
range, and several new studies consider the role of DOC in controlling pH and
subsequent effects on biota. In the 2008 ISA and in new research, few or no fish species
are found in lakes and streams that have very low ANC (near zero; Figure 8-4 and
Table 8-3) and low pH (near 5.0). The number of fish species generally increases at
higher ANC and pH values. A1 is very toxic to fish, and thresholds to elevated
concentrations of this metal in acidified waters are summarized in Table 8-4.
1.6.3.2 National-Scale Sensitivity, Biological Recovery, and Critical Loads
The extent and distribution of acid-sensitive freshwater ecosystems and sensitive regions
in the U.S. were well known at the time of the 2008 ISA. Measured data on lake and
stream ANC across the U.S. exhibit clear spatial patterns (Figure 8-11). Surface waters in
the U.S. that are most sensitive to acidification are largely found in the Northeast,
southern Appalachian Mountains, Florida, the upper Midwest, and the mountainous West
(Chapter 1, Figure 1-9). Levels of acidifying deposition in the West are low in most
June 2018
72
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
areas, acidic surface waters rare, and the extent of chronic surface water acidification to
date has been very limited. However, episodic acidification occurs in both the East and
West at sensitive locations, and this is partly natural and partly caused by humans.
Geographic patterns in acidification sensitivity vary in response to spatial differences in
geology, hydrologic flow paths, presence and depth of glacial till, climate, and other
factors (Appendix 8.5. IV In the eastern U.S., acid-sensitive ecosystems are generally
located in upland, mountainous terrain underlain by weathering-resistant bedrock. Some
of the most in-depth studies of the effects of acid stress on fish were conducted in streams
in Shenandoah National Park in Virginia and in lakes in the Adirondack Mountains of
New York. Effects on fish have also been documented in acid-sensitive streams of the
Catskill Mountains of southeastern New York, and the Appalachian Mountains from
Pennsylvania to Tennessee and South Carolina.
73
DRAFT: Do Not Cite or Quote

-------
Surface Water Critical Loads for Acidity
(2) Negative Critical Loads = 0.1 meq/m2/yr
	States	(3) NCLD v2.5 - 3/17/2015
ANC = acid-neutralizing capacity; meq = milliequivalent yr = year.
Source: http://nadp.sws.uiuc.edu/committees/clad/db/NCLDMapSummary 2015.pdf.
Figure 1-9 Surface water critical loads for acidity in the U.S. 10th percentile
aggregation for 36-km2 grids with S and N.
1
2	Biological recover} in acid-affected areas is discussed in Chapter 1.11. Typically
3	biological recovery occurs only if chemical recovery (Appendix 7.1.5.1) is sufficient to
4	allow growth, survival, and reproduction of acid-sensitive plants and animals. Surface
5	water chemistry recovery varies by region with the strongest evidence for improvement
6	in the Northeast and little or no recovery in central Appalachian streams. Acidification
7	and recovery of fresh waters will also be affected by the physical, chemical, and
8	biological modifications to acid inputs projected to occur with changes in annual mean
9	temperature and magnitude of precipitation (Appendix 8.5.3). As reported m the 2008
10	ISA and in new studies, biological recovery lags behind chemical recovery in many
June 2018
74
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
systems (see Appendix 8.4). The time required for biological recovery is unknown and
only partial biological recovery may be possible.
Since the 2008 ISA, considerable CL research has focused on aquatic acidification in the
U.S. The CLs for deposition are expressed in eq/ha/yr of S, N, or S + N because one or
both pollutants could be contributing to the observed effects. New empirical CLs include
571 eq N/ha/yr in the Northeast and 286 eq N/ha/yr in the West to prevent episodic
acidification in high-elevation lakes and to maintain an ANC of 74 eq/ha/yr in
high-elevation lakes of the Sierra Nevada (Table 8-7). Steady-state CLs have been
derived at many locations since the 2008 ISA (Table 8-8). Steady-state CLs of acidifying
deposition for lakes in the Adirondack Mountains (1,620 eq/ha/yr) and for the central
Appalachian streams (3,700 eq/ha/yr) were calculated to maintain a surface water ANC
of 50 |icq/L on an annual basis (NAPAP. 2011). CL values of less than 500 eq/ha/yr were
calculated for one-third of streams in the Blue Ridge ecoregion, to maintain stream ANC
at 50 (ieq/L. For lakes in Class I and II wilderness areas in the Sierra Nevada, CLs for
acidifying deposition in 2008 were estimated at ANC values of 0, 5, 10, and 20 |icq/L.
which span the range of minimum ANC values observed in Sierra Nevada lakes. The
median CL for granitic watersheds based on a critical ANC limit of 10 j^icq/L was
149 eq/ha/yr. Slightly more than one-third of these lakes had estimated rates of acidifying
deposition higher than their CL.
In addition to the steady-state and empirical loads described above, CL estimates have
been derived from dynamic modeling (Appendix 8.5.4). For example, there is new work
on simulated past and future effects of N and S on stream chemistry in the Appalachians
and Adirondack Mountain lakes. In 12 watersheds in the Great Smoky Mountain
National Park, target levels of ANC to protect aquatic life were used and ranged from
minimal (0 j^ieq/L) to considerable protection (50 j^ieq/L). For the 12 study streams, target
levels ofNCV + SO42 deposition ranged from 270 to 3,370 eq/ha/yr to reach an ANC of
0 (j,eq/L by 2050 and 0-1,400 eq/ha/yr to reach an ANC of 50 |icq/L by 2050. However,
the majority of streams could not achieve the ANC target of 50 (ieq/L. Modeling also
suggests that complete recovery from acidification may not be possible by the Year 2100
at all sites in the southern Blue Ridge region (Sullivan et al.. 201 lb) even if S emissions
cease entirely. In Shenandoah National Park, MAGIC modeling based on simulations of
14 streams identified a target load of about 188 eq kg S/ha/yr to achieve an
ANC = 50 (ieq/L (preindustrial level based on hindcast simulations) in 2100 in sensitive
streams. In a dynamic modeling simulation in the Adirondacks Mountains, about 30% of
the lakes in the region had a target load <500 eq/ha/yr to protect lake ANC to 50 j^icq/L
(Sullivan et al.. 2012a). Future decreases in SO42 deposition are suggested to be more
effective in that region in increasing Adirondack lake water ANC than equivalent
decreases in NO3 deposition. In another modeling study of 20 Adirondack watersheds,
75
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
estimates of preindustrial ANC for the study lakes ranged from 18 to 190 (ieq/L, and
simulations estimate that lake ANC have decreased by 26 to 100 j^icq/L as a legacy of
acidification.
1.7 Estuarine and Near-Coastal Ecosystem Nitrogen Enrichment
and Acidification
For estuaries (areas where fresh water from rivers meets the salt water of oceans) and
near-coastal systems, causal determinations from the 2008 ISA are further supported and
strengthened by additional studies, and there are two new causal statements on the
emerging topic of nutrient-enhanced coastal acidification (Chapter 1, Table 1-1).
Estuaries support a large biodiversity of flora and fauna and play a role in nutrient
cycling. N from atmospheric and other sources contributes to increased primary
productivity, leading to eutrophication (Figure 10-1). and N pollution is the major cause
of harm to the majority of estuaries in the U.S. (Appendix 10). Source apportionment
data in the 2008 ISA and newer studies indicate that atmospheric contributions to
estuarine N are heterogeneous across the U.S., ranging from <10% to approximately 70%
of total estuary N inputs (Table 7-6). In estuaries, increasing nutrient over-enrichment
leading to eutrophication is indicated by water quality deterioration, resulting in
numerous harmful effects, including areas of low dissolved oxygen (DO) concentration
(hypoxic zones), species mortality, and HABs. New studies support the 2008 ISA's
causal findings that increased N loading to coastal areas can alter biogeochemical
processes and lead to shifts in community composition, reduced biodiversity, and
mortality of biota. The current causal statement of biological effects of N enrichment in
estuarine ecosystems has been expanded to include total primary production, altered
growth, and total algal community biomass (Chapter 1, Table 1-1). Since the 2008 ISA, it
has been suggested that, in addition to atmospheric sources of CO2, nutrient-enhanced
productivity may contribute to acidification of coastal waters. New causal determinations
have been added to the current ISA on N deposition and increased nutrient enhanced
coastal acidification and associated biological effects (Chapter 1, Table 1-1). Increased
ocean acidification interferes with the ability of some organisms to build shells, although
the contribution of nutrient-enhanced coastal acidification is not yet understood.
1.7.1 Estuary and Near-Coastal Biogeochemistry
In the 2008 ISA, the evidence was sufficient to infer a causal relationship between
reactive N deposition and biogeochemical cycling of N and C in estuarine and
near-coastal marine systems. Evidence reviewed in the 2008 ISA, along with new studies
June 2018
76
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
indicate elevated N inputs to coastal areas can alter key processes that influence N and C
cycling in near-coastal environments. As N fluxes to coastal areas have increased in
recent decades in many parts of the U.S., the varying rates of different N cycling
processes within estuaries themselves can also affect the magnitude of eutrophication
experienced as a result of external N enrichment. Nitrogen additions not only cause the
total pool of N to be larger, but may also perturb N cycling in such a way that the system
may exacerbate eutrophication to a greater extent than expected based on N additions
alone. Research conducted since the 2008 ISA has shown that many of these N cycling
processes are more important in the estuarine environment than previously understood.
The removal of N through denitrification is a valuable ecosystem service in terms of
constraining the extent and magnitude of eutrophication. Additional research has
established dissimilatory NOs reduction to NH4 (DNRA) as a more important N
reduction pathway in some estuaries. Ammonium produced via DNRA can lead to
enhanced productivity and respiration, which may exacerbate hypoxia. Recent studies
indicate that DNRA rates are higher in warmer months and can also take up a larger
percentage of total N reduction activity when temperatures are higher. The roles of
sedimentary microbial processes of denitrification and anammox have been further
characterized. New research has shown that the community ofN fixing microorganisms
is more diverse in estuarine and coastal waters than previously thought, and that N
fixation occurs more widely than previously assumed. Influence of benthic macrofauna
on N cycling has received increased research attention in part due to the potential for
these organisms to mitigate external N enrichment. New research further supports
conclusions of the 2008 ISA, and the body of evidence is sufficient to infer a causal
relationship between N deposition and the alteration of biogeochemistry in estuarine
and near-coastal marine systems.
Since the 2008 ISA, a number of papers have identified links between nutrient
enrichment and acidification of coastal waters (Appendix 7.2.4). One of the initial studies
found that CO2 production during decomposition of organic matter delivered to coastal
zones from rivers experiencing eutrophication has enhanced the acidification of coastal
subsurface waters in the Gulf of Mexico and the East China Sea (Cai etal.. 2011c) and
additional studies provide evidence of acidification in estuaries due to this mechanism
(Appendix 7.2.7.2). Additional CO2 in the water column is produced from respiration of
living algae and seagrasses. The CO2 produced in eutrophic estuarine waters combines
with water molecules, producing carbonic acid, which makes the water more acidic.
Nutrient-enhanced coastal acidification tends to occur in locations where there is either
thermal or saline stratification. Modeling of coastal acidification via N enrichment and
atmospheric CO2 dissolution suggests that the combined effects of these two pathways
are synergistic. The body of evidence is sufficient to infer a likely causal relationship
between N deposition and increased nutrient-enhanced coastal acidification.
77
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
1.7.1.1
Nitrogen Enrichment
Estuarine biogeochemistry is complicated because it directly controls more than just the
N cycle; the response to N loading resulting in eutrophication impacts the chemical
cycling of metals and DO (Appendix 1.23). redox conditions, pH (Appendix 7.2.4). and
ultimately energy transfer (e.g., food webs from microbes to humans). The response to N
loading is also tightly controlled by the availability of organic matter (i.e., C) and its
lability and reactivity. Excess nutrient inputs are occurring within the context of other
stressors such as climate change (Appendix 7.2.6.12') and rising atmospheric CO2, which
further modify coastal biogeochemistry (Doncv. 2010). As reported in the 2008 ISA,
estuaries are generally N limited, and have received sufficiently high levels of N input
from human activities (including deposition, agricultural runoff, and wastewater) to cause
eutrophication. Highly variable environments within estuaries are characterized by a
gradient of increasing salinity toward the ocean. As N moves downstream, some fraction
is taken up by phytoplankton or removed by microbial denitrification. Key processes that
influence N cycling include hypoxia, nitrification, denitrification, and decomposition.
Until recently, it was generally believed that NH3 oxidation was accomplished only by
Proteobacteria in marine environments. New research discovered that some archaea can
also oxidize NH3. These ammonia-oxidizing archaea are dominant in some estuaries,
while ammonia-oxidizing bacteria are more important in others.
In the complex environment of the freshwater-to-ocean continuum, there are many
chemical and biological indicators of eutrophic condition. One approach is to measure
total nutrient loading and concentrations; however, these data need to be interpreted in
the context of the physical and hydrological characteristics that determine ecosystem
response. Water quality measures such as pH and DO, along with key biological
indicators such as chlorophyll a, phytoplankton abundance, HABs, macroalgal
abundance, and submerged aquatic vegetation (SAV; rooted vascular plants that do not
emerge above the water), can all be used to assess responses to nutrient loading
(Table 10-1). Nitrogen removal from the estuary is also influenced by faunal as well as
microbial communities.
Organic particles in coastal regions sink to the sediment-water interface where they
accumulate and decompose. Decomposition of these organic particles transforms
nutrients and depletes O2 in the water. Decreasing DO can create hypoxic (<2 mg/L of
dissolved O2) or anoxic zones inimical to fish and other aerobic life forms. Oxygen
depletion largely occurs only in bottom waters under stratified conditions, not throughout
the entire water column. This can result in seasonal hypoxia in shallow coastal regions,
particularly those that receiving high inputs of nutrients from coastal rivers. Development
of hypoxia is increasingly a concern in estuaries across the U.S. (Appendix 10.2.4V
June 2018
78
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
Since the 2008 ISA, N enrichment has been recognized as a potential contributing factor
to acidification of coastal waters (Appendix 10.5). Dissolution of atmospheric
anthropogenic CO2 into the ocean has led to long-term decreases in pH. With increasing
N inputs to coastal waters, decomposition of excess organic matter associated with
eutrophication adds CO2 to the water column (Sunda and Cai. 2012; Cai etal.. 2011c;
Howarth et al.. 201IV Models show that while the impact of each acidification pathway
(N enrichment or atmospheric CO2 dissolution) may be moderate, the combined effect
may be much larger than would be expected from the additive effects of each pathway
(Sunda and Cai. 2012; Cai et al.. 2011c').
1.7.1.2 Models
Since the 2008 ISA, several new applications of existing models have quantified
eutrophication processes in estuaries and near-coastal marine ecosystems. These have
included studies that focused primarily on N cycling or hypoxia. Other models of
estuarine eutrophication focus on N load apportionment, or on relationships between N
loads and ecological endpoints. Since the 2008 ISA, SPARROW has been used to
estimate total N loads within watersheds to determine sources of N to streams and rivers;
it has also been applied at regional and national scales. Additional models and tools that
include the contribution of N directly from the atmosphere have been applied to U.S.
estuaries, including the Watershed N Loading Model (NLM) and the Watershed
Deposition Tool (WDT). The latter was developed by the U.S. EPA to map atmospheric
deposition estimates to watersheds using wet and dry deposition data from CMAQ
(Schwedc et al.. 2009). This tool links air and water quality modeling data for use in total
maximum daily load (TMDL) determinations and analysis of nonpoint-source impacts.
New model applications include studies that focused primarily on endpoints of N cycling,
hypoxia, and HABs. Models of coastal eutrophication are described in greater detail in
Appendix 7.2.8.
1.7.1.3 National-Scale Sensitivity
Sensitivity of estuaries to biogeochemical changes associated with N enrichment varies
across the U.S. The biogeochemical sensitivity of estuaries and near coastal areas will be
discussed together with biological sensitivity to N enrichment in Chapter 1.7.4.
June 2018
79
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
1.7.2
Biological Effects of Nitrogen Enrichment
In the 2008 ISA, evidence was sufficient to infer a causal relationship between N
deposition and the alteration of species richness, community composition, and
biodiversity in estuarine ecosystems. The strongest evidence for a causal relationship was
from changes in biological indicators of nutrient enrichment [chlorophyll a, macroalgal
(seaweed) abundance, HABs, DO, and changes in SAV (Table 10-1)1. Some indicators,
such as chlorophyll a, are directly linked to nutrient enrichment and provide evidence of
early ecosystem response; other indicators, such as low DO and decreases in SAV,
indicate more advanced eutrophication. Phytoplankton are the base of the coastal food
web and increases in primary producer biomass and altered community composition
associated with increased N can lead to a cascade of direct and indirect effects at higher
trophic levels. At the time of the 2008 ISA, N was recognized as the major cause of harm
to the majority of estuaries in the U.S. Since 2008, new paleontological studies,
observational studies, and experiments have further characterized the effects of N on
phytoplankton growth and community dynamics, macroinvertebrate response, and other
indices of biodiversity. For this ISA, new information is consistent with the 2008 ISA and
the causal determination has been updated to reflect more specific categories of effects to
include total primary production, altered growth and total algal community biomass. This
new research strengthens the evidence base and is consistent with the 2008 ISA
(Chapter 1, Table 1-1) that the body of evidence is sufficient to infer a causal
relationship between N deposition and changes in biota including total primary
production, altered growth, total algal community biomass, species richness,
community composition, and biodiversity due to N enrichment in estuarine
environments.
Since the 2008 ISA, additional evidence has shown that reduced forms of atmospheric N
play an increasingly important role in estuarine and coastal eutrophication and HAB
dynamics. New studies emphasize that N inputs interact with physical and hydrologic
factors to increase primary productivity and eutrophication in coastal areas.
Climate-related changes including temperature, precipitation, wind patterns, extreme
weather events, stronger estuary stratification, increased metabolism and organic
production, and sea-level rise are all expected to modify coastal habitats
(Appendix 10.1.4.1).
1.7.2.1 Primary Producers
Algae are the base of the coastal food web, and the 2008 ISA showed that changes in
chemical composition of N inputs can shift the algal community and cascade up the food
June 2018
80
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
web. Chlorophyll a is broadly recognized indicator of phytoplankton biomass and is used
as a proxy for assessing effects of estuarine nutrient enrichment. It can signal an early
stage of water quality degradation related to nutrient loading and is incorporated into
water quality monitoring programs and national-scale assessments including U.S. EPA's
National Coastal Condition Assessment (Appendix 7.2.7). Phytoplankton sampling,
microcosms studies and sediment core analysis have shown changes in phytoplankton
community structure in estuaries with elevated N inputs (Appendix 10.3). These shifts at
the base of the food web to species that are not as readily grazed (e.g., cyanobacteria,
dinoflagellates) have a cascade of effects including poor trophic transfer and an increase
in unconsumed algal biomass, which could stimulate decomposition, O2 consumption,
and the potential for hypoxia.
There is consistent and coherent evidence that the incidence of HAB outbreaks is
increasing in both freshwater and coastal areas, a problem that has been recognized for
several decades (Appendix 10.2.2). Of the 81 estuary systems for which data were
available for the National Estuarine Eutrophication Assessment (NEEA), 26 exhibited a
moderate or high symptom expression for nuisance or toxic algae (Bricker et al.. 2007).
Since the 2008 ISA, HAB bloom conditions and effects of HAB toxins on wildlife have
been further characterized (Appendix 10.2.2). Release of toxins during HABs can be
harmful to fish and shellfish, and these toxins may be transferred to higher trophic levels.
The form of N affects phytoplankton growth and toxin production of some HAB species.
Increasing loads of NH3+/NH4+ have been linked to the expansion of HABs and altered
phytoplankton community dynamics (Appendix 10.3.3). Cyanobacteria, and many
chlorophytes and dinoflagellates may be better adapted to the use of NH44" while diatoms
generally thrive in oxidized forms of N such as NO3 (Figure 10-7).
Macroalgal (seaweed) growth is also stimulated by increased N inputs, which increase
the dominance of faster growing benthic or pelagic macroalgae to the exclusion of other
species (Appendix 10.2.3). Studies published since the 2008 ISA provide further
evidence that macroalgae respond to the form of N with some species showing greater
assimilation and growth rates with NH44" than with NO3 . Increased abundance of
macroalgae, which block light, and increased epiphyte loads on the surface of SAV may
reduce the growth and biomass of SAV. SAV including the eelgrass Zostera marina are
important ecological communities found within some coastal bays and estuaries that are
sensitive to elevated nutrient loading and loss of this habitat can lead to a cascade of
ecological effects because many organisms are dependent upon seagrasses for cover,
breeding, and as nursery grounds. Recently, presence of seagrass beds was linked to
decreased bacterial pathogens of humans, fishes, and invertebrates in the water column
and lower incidence of disease in adjacent coral reefs (Appendix 10.2.5). The 2008 ISA
reported correlations between increased N loading and declines in SAV abundance and
81
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
newer studies have further characterized this relationship. In a survey of southern New
England estuaries, reduced eelgrass extent was observed at increased watershed N
loading. New studies have characterized the role of invertebrate mesograzers, such as
small crustaceans and gastropods, in controlling algal growth, potentially buffering
eutrophication effects on seagrass communities (Appendix 10.3.7). Macroalgae may not
be a good indicator of eutrophication in some upwelling-influenced estuaries in the
Pacific Northwest as an increase in macroalgal biomass in these systems does not appear
to be associated with temporal declines in eelgrass (Appendix 10.2.3).
1.7.2.2 Bacteria and Archaea
Ammonia-oxidizing prokaryotes carry out nitrification in estuarine waters.
Ammonia-oxidizing archaea are relatively recently described, and several studies since
the 2008 ISA have considered community responses of ammonia-oxidizing bacteria and
ammonia-oxidizing archaea. Community structure of ammonia-oxidizers is related to
nutrient inputs and affected by the form of available N (Appendix 10.3.4V
1.7.2.3 Invertebrates
The community of benthic organisms shifts toward shorter life spans and smaller body
size in coastal areas with severe seasonal hypoxia (Appendix 10.2.4). Reduced species
density and diversity in the northern Gulf of Mexico are linked to persistent hypoxic
events. The form of N present has been shown to affect molluscan taxonomic
assemblages (Appendix 10.3.5). Shifts in algal composition and productivity can affect
growth of shellfish that feed on phytoplankton. Shellfish contribute to N and C cycling
and can improve water quality, and recent research has explored the use of these
organisms for coastal N remediation (Appendix 7.2.6.11). Harvest of shellfish for human
consumption removes nutrients from estuaries.
N enrichment is one of several factors linked to increased disease susceptibility,
bleaching, and reduced calcification rate in corals (Appendix 10.4.2). Several studies
have isolated effects of N, which affects corals via pathways that are distinct from P. The
threatened status of staghorn coral (Acropora cervicornis) and elkhorn coral (Acropora
palmata) under the U.S. Endangered Species Act has been linked to indirect N pollution
effects, specifically low DO, algal blooms that alter habitat, and other non-nutrient
stressors (Hernandez et al.. 2016).
June 2018
82
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
1.7.2.4 Fish
Fish biodiversity is altered by increased N inputs and resulting changes in biological and
chemical indicators (Appendix 10.3.6). Many fish are unable to persist at DO levels
below 2 mg/L (Figure 10-4). Recent studies in the southern Gulf of Saint Lawrence have
linked SAV loss to declines in fish biodiversity, although organisms did not change
positions within food webs. In laboratory conditions, turbidity associated with
eutrophication alters fish reproductive behaviors. Hypoxia has also recently been shown
to affect reproduction in fish. For example, hypoxia acts as an endocrine disruptor in
Atlantic croaker (Micropogonias undulatus; Appendix 10.2.4).
1.7.3 Biological Effects of Nutrient-Enhanced Coastal Acidification
Coastal acidification (Chapter 1.7.1). which can be exacerbated by elevated N input, is
projected to alter marine habitat, have a wide range of effects at the population and
community level, and impact food web processes. Newer studies show that organisms
that produce calcium carbonate shells are impacted by increasing acidification of ocean
waters (Appendix 10.5). Decreased concentration of carbonate ions (which organisms
such as calcareous plankton, oysters, clams, sea urchins, and coral take up to build shells)
are observed in acidic conditions. With increasing N inputs to coastal waters CC^in the
water column is produced from degradation of excess organic matter from changing land
use, as well as respiration of living algae and seagrasses, which in turn can make the
water more acidic. Documented declines of oyster production on the U.S. west coast are
linked to ocean acidification. Research on costressors associated with conditions of
coastal acidification and eutrophication suggest that interactions between elevated CO2,
decreasing pH, and nutrient inputs are complex. The body of evidence is suggestive of,
but not sufficient to infer, a causal relationship between N deposition and changes in
biota including altered physiology, species richness, community composition, and
biodiversity due to nutrient-enhanced coastal acidification.
1.7.4 National-Scale Sensitivity and Critical Loads
The NEEA, the most recent comprehensive survey of eutrophic conditions in U.S.
estuaries conducted by the National Oceanic and Atmospheric Administration, defined
eutrophication susceptibility as the natural tendency of an estuary to retain or flush
nutrients (Bricker et al.. 2007). In estuaries that have longer water residence times,
nutrients are more likely to lead to eutrophic conditions (Appendix 10.1.4). As reported
in the 2008 ISA and newer studies, nutrient loading accelerates hypoxia, which is more
June 2018
83
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
likely in marine waters with limited water exchange, water column stratification, and
high production and settling of C to bottom waters. Other factors identified in the
2008 ISA that increase estuary sensitivity to eutrophication include human population,
agricultural production, and the size of the estuary relative to its drainage basin. The
NEEA reported that the most eutrophic estuaries in the U.S. occur in the mid-Atlantic
region, and the estuaries with the lowest degree of eutrophication are in the North
Atlantic (Figure 10-2). Estuaries identified in the 2008 ISA as susceptible to
eutrophication include the Chesapeake Bay, Pamlico Estuary in North Carolina, Long
Island Sound, as well as along the continental shelf adjacent to the Mississippi and the
Atchafalaya River discharges to the Gulf of Mexico. New research at the regional scale
includes long-term studies of several coastal systems on trends in coastal water quality
and chemistry. A 23-year study of the Chesapeake Bay concluded that water quality has
decreased and chlorophyll a levels have increased since 1986, in part due to long-term
climate trends (see Appendix 10.2.5).
Since the 2008 ISA, there is additional information on the extent and severity of
eutrophication and hypoxia in sensitive regions. Areas of eutrophication-related hypoxia
are found on the U.S. east and west coasts and the Gulf of Mexico (Figure 10-5). The
2008 ISA reported that the largest zone of hypoxic coastal water in the U.S. was the
northern Gulf of Mexico on the Louisiana-Texas continental shelf. In the summer of
2017, the hypoxic zone in the Gulf was the largest ever measured at 14,123 km2
(8,776 mi2) (U.S. EPA. 2017e). Atmospheric deposition to watersheds in the
Mississippi/Atchafalaya River Basin contributes approximately 16 to 26% of the total N
load to the Gulf of Mexico (Appendix 10.2.4). Long Island Sound also experiences
periods of anoxia. In other U.S. coastal systems, hypoxia incidence is increasing, but DO
impacts are relatively limited temporally and spatially. In the Pacific Northwest, coastal
upwelling can be a large source of nutrient loads and advection of upwelled water can
introduce hypoxic water into estuaries that is not related to anthropogenic sources.
The NEEA suggested that only a small fraction of the estuary systems evaluated reported
moderate to high SAV loss (Bricker et al.. 2007). mostly in the mid-Atlantic region.
While seagrass coverage is improving in some estuaries, such as Tampa Bay (Tampa Bay
Case Study, Appendix 16). many estuaries continue to see declines in seagrass extent.
SAV is often at a competitive disadvantage under N enriched conditions due to the fast
growth of opportunistic macroalgae that preferentially take up NH4 and can block light
from seagrass beds.
There are thresholds of response identified for some biological and chemical indicators of
N enrichment in estuaries (Appendix 10). Chlorophyll a is an indicator of phytoplankton
biomass, and thus, a proxy for assessing estuarine nutrient enrichment. In general,
84
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
0-5 jj.g/L chlorophyll a is considered to be good condition, concentrations between 5 and
20 (ig/L are classified as fair condition, and concentrations of >20 (ig/L indicate poor
conditions (Table 10-2). A new response threshold of tidal-averaged total N
concentration of <0.34 mg/L has been identified for healthy eelgrass in Massachusetts
waters. Markedly decreased eelgrass coverage is observed atN loading rates
>100 kg N/ha/yr, and levels above 50 kg N/ha/yr are likely to impact SAV habitat extent
in shallow New England estuaries (Table 10-4). Greaver et al. (2011) identified the range
of 50-100 kg N/ha/yr total N loading as the empirical critical load for loss of eelgrass
based on Latimer and Rego (2010). In terms of DO, concentrations of 0 mg/L are anoxic,
0-2 are indicative of hypoxic conditions, and 2-5 mg/L are biologically stressful
conditions (Figure 10-4). Oxygen depletion largely occurs only in bottom waters under
stratified conditions, not throughout the entire water column.
The indicators of nutrient enrichment in coastal areas (chlorophyll a, HABs, macroalgal
abundance, DO, SAV, and benthic diversity) have been incorporated into indices of
coastal eutrophication. In the 2008 ISA, the Assessment of Estuarine Tropic Status
(ASSETS) categorical Eutrophication Condition index (ECI) developed for the NEEA
was used as an assessment framework for coastal U.S. estuaries (Bricker et al.. 2007).
Additional indices of estuarine functioning that incorporate biological indicators have
since been developed both in the U.S. and internationally (Appendix 10.2.6).
Comparisons of these frameworks have identified robust methods to measure estuarine
response, such as incorporation of annual data, frequency of occurrence, spatial coverage,
secondary biological indicators, and a multicategory rating scale.
Since the 2008 ISA, N enrichment has been linked to coral bleaching and reduced
calcification rates (Appendix 10.4.2). Near-coastal coral reefs in the U.S. occur off south
Florida, Texas, Hawaii, and U.S. territories in the Caribbean and Pacific.
1.8 Wetland Ecosystem Nitrogen Enrichment and Acidification
New evidence, including new critical loads, supports and strengthens the causal findings
from the 2008 ISA regarding N enrichment effects in wetlands (Chapter 1, Table 1-1). In
freshwater and coastal wetland ecosystems, deposition of N and S do not tend to cause
acidification-related effects at levels currently common in the U.S IYU.S. EPA. 2008a). in
Annex B], However, the 2008 ISA documented that wetlands can be sensitive to N
enrichment and eutrophication effects. Newer studies have characterized N effects on
biogeochemistry, physiology, biodiversity, national sensitivity, and critical loads for
freshwater and coastal wetlands; coastal wetlands are typically tolerant of higher N
loading than freshwater wetlands.
June 2018
85
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
1.8.1
Wetland Biogeochemistry
In the 2008 ISA, evidence was sufficient to infer a causal relationship between N
deposition and the alteration of wetland biogeochemical cycling. Although sources and
rates of N inputs vary widely among wetlands, N deposition contributes substantially to
total loading in many wetlands. This additional N alters multiple aspects of
biogeochemistry, including C cycling, N cycling, and release of nutrients to
hydrologically connected surface waters. New research together with the information
included in the 2008 ISA shows that the body of evidence is sufficient to infer a causal
relationship between N deposition and the alteration of biogeochemical cycling in
wetlands.
The 2008 ISA reported that N enrichment altered N cycling in wetland ecosystems.
Chemical indicators of N deposition in wetlands include NOs and NH44" leaching, DON
leaching, N mineralization, denitrification rates, and N2O emissions. A wetland can act as
a source, sink, or transformer of atmospherically deposited N, and these functions vary
with season and hydrological conditions. Vegetation type, physiography, local hydrology,
and climate all influence source/sink N dynamics in wetlands. A new synthesis of global
wetland data showed that wetland reactive N removal and water quality improvement is
proportional to reactive N load, and removal efficiency is 26% higher in nontidal than
tidal wetlands, while a new meta-analysis shows that N enrichment increases wetland
N2O emissions by 207%. New studies have also evaluated the effects of N loading/N
addition on other endpoints related to N cycling in peat bog, riparian, mangrove, and salt
marsh wetlands (see Appendix 11.3.1). The endpoints evaluated include ecosystem N
retention, wetland export of N to surface waters, N fixation, N mineralization,
denitrification, emission of N2O, and bacterial abundance, activity, and composition in
wetland soils. The results of North American studies are summarized in Figure 11-2.
Across studies, N enrichment decreases the ability of wetlands to retain and store
nitrogen, which may diminish the wetland ecosystem service of improving water quality.
In the 2008 ISA, evidence from Canadian and European peatlands showed that N
deposition had negative effects on Sphagnum (moss) bulk density and mixed effects on
Sphagnum productivity depending on the history of deposition. There is new information
on how N deposition alters biogeochemical cycling of C in wetlands. Chemical indicators
of N deposition in wetlands include soil organic matter, total soil C or peat C, CO2
emissions, and CH4 emissions. Long-term C storage is an important ecosystem service of
wetlands for which measures of physical marsh stability can serve as a proxy, and
physical indicators of N deposition can include temperature, bulk density, physical
resistance, and soil water content. In addition, changes to plant growth rates and
productivity indicate altered C cycling in wetlands, and are summarized in Chapter 1.8.2.
June 2018
86
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
The literature evaluates the effects of N deposition, N loading, or experimental N
addition on C cycling in bogs, fens, riparian or intermittent marshes, freshwater tidal
marshes, mangroves, and salt marshes (see Appendix 11.3.2). Significant effects of N
loading upon biogeochemical cycling of C in North American wetlands (in which the N
addition was 500 kg N/ha/yr or lower) are summarized in Figure 11-3. N enrichment
decreases wetland retention of C, as indicated by new studies and a new meta-analysis
that show that N enrichment increases methane production in salt marshes. New studies
of marshes along the Gulf Coast and East Coast find that N enrichment also decreases the
bulk density of salt marshes, making marshes less resilient to physical stresses from tidal
or storm flooding, and may accelerate coastal marsh loss.
1.8.2 Biological Effects of Wetland Nitrogen Enrichment/Eutrophication
In the 2008 ISA, evidence was sufficient to infer a causal relationship between N
deposition and the alteration of species richness, species composition, and biodiversity in
wetlands. New evidence is presented in the following sections regarding the effects of N
upon wetland plant physiology, architecture, demography, and biodiversity. The body of
evidence is sufficient to infer a causal relationship between N deposition and the
alteration of growth and productivity, species physiology, species richness,
community composition, and biodiversity in wetlands.
1.8.2.1 Growth, Productivity, and Physiology
In the 2008 ISA, evidence from Canadian and European bogs and fens showed that N
deposition had negative or mixed effects on Sphagnum (moss) productivity, depending on
history of deposition. In Canadian ombrotrophic peatlands experiencing deposition of
2.7-8.1 kg N/ha/yr, peat accumulation increased with N deposition, but accumulation
rates had slowed by 2004, indicating a degree of N saturation. Coastal wetlands
responded to N enrichment with increased primary production, shifting microbial and
plant communities and altering pore water chemistry, although many of the studies in
coastal wetlands used N enrichment levels more similar to those of wastewater than
atmospheric deposition. New research on N enrichment effects on growth and
productivity was conducted in ombrotrophic bogs, intermittent wetlands, freshwater tidal
marsh, mangroves, and coastal salt marshes (see Appendix 11.4). Ecological endpoints
evaluated to assess N loading effects on growth and productivity include plant
aboveground biomass and productivity, plant belowground biomass of roots and
rhizomes, and growth rates, and are summarized along with N effects on C cycling in
Figure 11-3. The effects of N additions on plant physiology were not addressed in the
June 2018
87
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
2008 ISA, but information regarding these effects is available for bogs and fens, riparian
wetlands, freshwater tidal marsh, mangroves, and salt marshes (see Appendix 11.5).
Ecological endpoints evaluated to assess N loading effects on plant physiology include
stoichiometry (i.e., nutrient concentrations and ratios of multiple nutrients in plant tissue),
nutrient acquisition efficiency (including insectivory rates in insectivorous plants),
nutrient use efficiency, and nutrient reabsorption efficiency, and are summarized in
Figure 11-4.
In general, across types of wetlands, nitrogen loading increases aboveground growth and
productivity while decreasing or not affecting belowground growth and productivity. In
bogs and fens, N deposition decreases growth of state-listed Sarracenia purpurea (purple
pitcher plant), and N enrichment increases aboveground productivity of emergent sedges
more than of peat-building moss species. These changes cascade up to affect biodiversity
in bogs and fens (see below, Chapter 1.8.2.2). In freshwater and tidal marshes, N
enrichment increases aboveground productivity while decreasing belowground
productivity, and this shift from belowground to aboveground plant productivity may
account for changes in wetland C storage (see Chapter 1.8.1).
Changes to plant physiology and stoichiometry vary by species tolerance to N and N
acquisition strategies. In bogs, N enrichment typically causes increased plant tissue N
concentrations, decreased N use efficiency, and decreased N resorption efficiency during
senescence. After several years of exposure to high rates of N loading, bog plants may
experience leaf N saturation and limitation by other nutrients (e.g., P, K, and Ca,
indicated by increasing reabsorption efficiencies), resulting in leaf damage in sensitive
species. S. purpurea (purple pitcher plant) decreases its dependence upon insectivory for
nutrition at N deposition rates of 4.4 kg N/ha/yr. In freshwater marshes, N enrichment
also increases plant tissue N concentrations while increasing P limitation and altering
resorption efficiencies.
Plant architecture was not addressed in the 2008 ISA, and demography was addressed
only for bogs and fens. Aboveground, plant architecture includes branching patterns, as
well as the size, shape, and position of leaves and flower organs. New studies find N
enrichment affects plant architecture in a salt marsh, in mangroves, in freshwater tidal
marshes, and in a riparian wetland (Appendix 11.6). In terms of plant demography, the
2008 ISA found positive population growth rates for S. purpurea at 0 or 1.4 kg N/ha/yr,
but population losses at 14 kg N/ha/yr, and that N deposition above 6.8 kg N/ha/yr
increases population extinction risk of S. purpurea. New studies show that N addition has
species-specific effects upon reproduction of West Coast salt marsh plant species, and
increases mortality across the global distribution of mangrove species (Appendix 11.7).
88
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
1.8.2.2 Biodiversity
In the 2008 ISA, evidence was sufficient to infer a causal relationship between N
deposition and the alteration of species richness, species composition, and biodiversity in
wetlands. Notably, the 2008 ISA cited 4,200 native plant species in U.S. wetlands, 121 of
which are federally endangered. Given their relative area, wetlands provide habitat to a
disproportionally high number of rare plants. Many wetland species have adapted to N
limited conditions, including endangered species in the genera Isoetes (3 endangered
species) and Sphagnum (15 endangered species), as well as insectivorous plants such as
pitcher plants (Sarracenia spp.) and sundews (Drosera rotundifolia).
Coastal wetlands responded to N enrichment with increased primary production,
changing microbial and plant communities, and altered pore water chemistry, although
many of the studies available in 2008 used high N enrichment levels more similar to N
loading from wastewater than from atmospheric deposition. New research since 2008
including N deposition gradient studies, experimental N addition studies, and
observational studies that show that N enrichment altered biodiversity in bogs and fens,
intermittent wetlands, freshwater wetlands, freshwater tidal wetlands, and coastal salt
marshes (see Appendix 11.8).
New research from wetland ecosystems strengthens the 2008 causal statement. New
research confirms that, as in terrestrial systems, N addition can decrease the abundance
and richness of sensitive species while increasing the abundance and richness of tolerant
species. In bogs and fens, N enrichment decreases the survival of insectivorous plants and
the cover of mosses, while increasing the cover of shrub species. In freshwater marshes,
N enrichment changes plant community composition, increases the abundance of and
stresses caused by invasive plant species, promotes the harmful algal species that produce
the toxin microcystin, and increases mosquito larvae that are vectors for zoonotic
diseases (see Figure 11-1). In freshwater tidal and coastal marshes, N enrichment changes
plant community composition, increases cover of invasive plant species, increases
herbivory by invertebrates, and increases herbivory by the invasive mammal Myocastor
coypus (nutria).
1.8.2.3 National Sensitivity and Critical Loads for Wetlands.
Freshwater and coastal wetlands tend to have different sensitivity to added N. Broadly,
wetlands that receive a larger fraction of their total water budget in the form of
precipitation are more sensitive to the effects of N deposition. For example, bogs
(70-100% of hydrological input from rainfall) are more sensitive to N deposition than
June 2018
89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
fens (55-83% as rainfall), which are more sensitive than coastal wetlands (10-20% as
rainfall).
Since the 2008 ISA, a CL for U.S. coastal wetlands has been established. The CL is based
on several different ecological endpoints, including plant community composition,
microbial activity, and biogeochemistry (63-400 kg N/ha/yr). Figure 11-6 shows a
comparison of the N CL for coastal wetlands with recent studies of ecological impacts of
N (at N levels of 100-250 kg N/ha/yr).
Since the 2008 ISA, two N critical loads (CLs) for U.S. freshwater wetlands have been
established. The CL for wetland C cycling, quantified as altered peat accumulation and
NPP, is between 2.7 and 13 kg N/ha/yr. The upper end of this critical load range is based
on measurements of wet deposition only (10 to 13 kg N/ha/yr), and therefore, does not
reflect total N loading. There is also a CL to protect biodiversity based on morphology
and population dynamics of the purple pitcher plant (Sarracenia purpurea) between
6.8-14 kg N/ha/yr. A more recent study across an N deposition gradient suggests that
purple pitcher plant populations experience negative effects of N deposition at rates lower
than this CL, but the more recent research has not yet been incorporated into the CL
framework. A comparison of freshwater wetland CLs to observed ecological impacts of
N from recent studies (4.4-500 kg N/ha/yr) is given in Figure 11-7.
1.9 Freshwater and Wetland Ecosystem Sulfur Enrichment
New evidence from wetland and freshwater aquatic ecosystems strengthens and extends
the causal findings of the 2008 ISA regarding nonacidifying sulfur effects, and provides
the basis for a new causal determination (Chapter 1, Table 1-1). New research together
with the information included in the 2008 ISA shows that the evidence is sufficient to
infer a causal relationship between S deposition and the alteration of Hg methylation in
surface water, sediment, and soils in wetland and freshwater ecosystems. New evidence
is sufficient to infer a new causal relationship between S deposition and changes in biota
due to sulfide phytotoxicity, including alteration of growth and productivity, species
physiology, species richness, community composition, and biodiversity in wetland and
freshwater ecosystems.
SOx deposition can have chemical and biological effects other than acidification,
particularly in flooded wetland soils and aquatic ecosystems. The 2008 ISA described
qualitative relationships between sulfate deposition and a number of ecological
endpoints, including altered S cycling, sulfide phytotoxicity, internal eutrophication of
aquatic systems, altered methane emissions, increased mercury (Hg) methylation, and
increased Hg loading in animals, particularly fish. Table 12-11 summarizes chemical
June 2018
90
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
concentrations that alter ecological endpoints and the quantitative relationships
describing effects of sulfate deposition. Recent research supports these relationships
between S deposition and ecological endpoints and provides the basis for SOx deposition
levels, water column sulfate concentrations, and water column sulfide concentrations
protective of plants and animals.
1.9.1 Biogeochemistry
SOx deposition alters biogeochemical processes via S enrichment including S cycling
(see Appendix 12.2.1). P cycling (see Appendix 12.2.4). C cycling (see
Appendix 12.2.5). and Hg cycling (see Appendix 12.3). The primary chemical indicator
for nonacidifying or enrichment effects of S in wetland and aquatic ecosystems is surface
water sulfate concentration, as it is for acidifying effects. The 2008 ISA reported that
chemical reduction of sulfate was an important indicator of SOx effects on water
chemistry, as the process generates ANC. There are no new studies on ANC generation
through sulfate reduction, although microbial sulfate reduction remains an active area of
research. In aquatic ecosystems for which atmospheric and terrestrial S inputs are similar
in magnitude to rates of microbial sulfate reduction, the products of microbial sulfate
transformation may be more reliable indicators of S enrichment effects than surface water
sulfate concentrations. These chemical indicators include methylmercury (MeHg),
sulfide, and phosphate.
MeHg is the most persistent and toxic form of Hg in the natural environment, and is
measured in surface water or aquatic sediments (MeHg concentration or the percentage of
MeHg in total Hg) to predict its effects upon biota. Several new studies demonstrate
significant positive relationships between surface water SO42 concentrations and water
or sediment MeHg concentrations (see Appendix 12.3.5). Another product of sulfate
reduction, sulfide (measured as surface water or sediment pore water S2 concentrations)
is also a water quality indicator of deposition effects upon biota. In freshwater
ecosystems with iron-rich sediments, sulfide may react with iron bound to phosphates in
the sediment to release phosphate into the water column, increasing primary productivity;
new research documents this process, referred to as internal eutrophication
(Appendix 12.2.4).
In terms of S enrichment effects upon carbon cycling, the 2008 ISA documented the
suppression of methane emissions in wetland soils by sulfate addition in several studies
and noted that 15 kg S/ha/yr suppressed methane emissions. Recent research has
confirmed that S enrichment increases the abundance or metabolic activity of
sulfate-reducing prokaryotes (SRPs), which under some conditions compete with
June 2018
91
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
methanogens and suppress their activity, in turn suppressing methane emissions
(Appendix 12.2.4V However, there are no new studies documenting S deposition effects
upon methane emissions in U.S. ecosystems.
1.9.2 Biological Effects of Sulfur Enrichment
Nonacidifying S effects upon biota include plant toxicity, changes in plant growth and
biodiversity, and increased Hg concentrations in biota. The toxicological effects of Hg
accumulation in animals were documented in the 2008 ISA and newer studies.
1.9.2.1 Sulfur Nutrient and Toxicity to Plants
Plants and other organisms require S as an essential nutrient. The deposition of S can
affect plant protein synthesis by affecting S availability for S containing amino acids,
which in turn will affect N uptake. The 2008 ISA documented the effects of sulfate
toxicity on plant development and reproduction at very high S loads. There is no new
evidence of S deposition effects upon plant S nutrition or sulfate toxicity. The product of
microbial sulfate reduction, sulfide, is an important plant toxin, and the 2008 ISA
documented sulfide phytotoxicity in European systems. Together with new research
showing sulfide phytotoxicity in North American wetlands, the body of evidence is
sufficient to infer a causal relationship between S deposition and changes in biota
due to sulfide phytotoxicity including alteration of growth and productivity, species
physiology, species richness, community composition, and biodiversity in wetland
and freshwater ecosystems.
The 2008 ISA showed that sulfide toxicity decreased biomass of wetland plants and
aquatic macrophytes in mesocosms under aquatic S concentrations higher than current
U.S. concentrations. In Europe, research showed that a threshold value of <48 mg
SC>427L in surface water would protect the sensitive aquatic species Stratiotes aloides and
Potamogeton acutifolius (not native to CONUS), as well as to protect Potamogeton
zosteriformis and Utricularia vulgaris, which are both native and widely distributed in
CONUS. New research has demonstrated sulfide phytotoxicity effects at current ambient
sulfide concentrations in multiple ecosystems within the U.S. (Appendix 12.2.3V Sulfide
decreased total plant cover and cover of dominant species in a New York fen, and
decreased the growth rate of Cladium jamaicense (sawgrass), a keystone species in the
Florida Everglades. Zizaniapalustris (wild rice) is an economically and culturally
important species sensitive to sulfide, and the Minnesota Pollution Control Agency has
developed a model for this species that calculates protective levels of water SOr
June 2018
92
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
concentrations, given iron and DOC concentrations in water bodies. A recent review
identifies sulfide thresholds between 0.3-29.5 mg S27L for altered growth, productivity,
physiology, or mortality of 16 freshwater wetland emergent plant and aquatic submerged
macrophyte species native to North America (see Table 12-2).
1.9.2.2 Sulfur Effects on Mercury Methylation
In the 2008 ISA, evidence was sufficient to infer a causal relationship between S
deposition and increased methylation of Hg, in aquatic environments where the value of
other factors is within adequate range for methylation. In the 2008 ISA, sulfur-reducing
bacteria (SRB) were identified as the organisms responsible for Hg methylation. New
evidence shows the ability to methylate Hg is more broadly distributed phylogenetically,
including both bacteria and archaea, which is why this document refers to
sulfate-reducing mercury methylators as SRPs rather than SRB (Appendix 12.3.2V In the
2008 ISA, wetland and lake-bottom sediments were identified as habitat for mercury
methylating SRPs. Recent research documents microbial mercury methylation in lakes, in
wetland sediments and moss, within periphyton, in marine ecosystems, and within
disturbed terrestrial forest soils (Appendix 12.3.2 and Appendix 12.3.3). Microbial
mercury methylation responsive to SOx deposition occurs in freshwater lakes, freshwater
wetlands, freshwater reservoirs, and freshwater agricultural areas (Appendix 12.3.4).
Between the 2008 ISA and new research, the body of evidence is sufficient to infer a
causal relationship between S deposition and the alteration of Hg methylation in
surface water, sediment, and soils in wetland and freshwater ecosystems.
Hg methylation is determined in part by surface water sulfate, as many strains of SRPs
possess the recently identified gene pair hgcAB, and pair their metabolism of C with both
dissimilatory sulfate reduction and mercury methylation (see Appendix 12.3.2 and
Figure 12-5). Microbial methylation rates are determined by other environmental
requirements of SRPs, including seasonality and temperature, pH, organic matter in water
and sediments, iron, nitrate, and salinity (Appendix 12.3.3). New research demonstrates
that Hg methylation occurs at ambient sulfate concentrations within U.S. water bodies.
Multiple lines of evidence support a relationship between sulfate surface water
concentrations and MeHg concentration or production in various freshwater systems.
Linear relationships between sulfate concentrations and MeHg concentrations were
observed in sediments of the South River, VA; across peat bogs in Minnesota and
Ontario; and across prairie pothole lakes in Saskatchewan (Figure 12-17). In addition to
the studies of lake and wetland sediments reviewed in the 2008 ISA, studies employing
lab incubations show that sulfate increases Hg methylation in samples from Adirondack
peat bogs, from South River, VA sediments, from periphyton growing in North American
June 2018
93
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
lakes and wetlands, and from leaf packs in Minnesota river water (Appendix 12.3.3.1).
Experimental addition of S to field mesocosms or whole ecosystems has shown that S
enrichment as wet S deposition increases MeHg in water, sediment, or biota, in Little
Rock Lake, WI; Bog Lake Fen, MN; the Experimental Lakes Area, Ontario; and the bog
experiment at Degero Stormyr, Sweden (Appendix 12.3.4.1). In observational studies of
S and Hg deposition, fish Hg concentrations decline with temporal declines in SOx
deposition in Isle Royale (a Class I area), fish Hg concentrations correlate positively with
Hg and S deposition across Texas ecoregions, and a 12-year study found fish Hg in
Voyageurs National Park (a Class I area) declined in lakes with decreasing S deposition
only when lake DOC remained constant (Appendix 12.3.5. IV New research is consistent
and coherent with the research presented in the ISA in demonstrating that sulfur
enrichment from SOx deposition stimulates mercury methylation in North American
ecosystems. Current research suggests that mercury methylation generally peaks between
10 and 100 mg SO42 /L in surface water, and quantitative relationships between S and Hg
such as target values or thresholds are reported in Table 12-12.
1.9.2.3 Sulfur, Mercury, and Animal Species
Mercury is a developmental, neurological, endocrine, and reproductive toxin across
animal species. The 2008 ISA documented Hg accumulation in fish, songbirds, four turtle
species, insectivorous passerine birds, and the common loon (Gavia immer). Recent
research also documented Hg accumulation in insectivore songbirds, bats, and fish in
agricultural wetlands. The 2008 ISA reported that 23 states had issued fish advisories by
2007 in response to the U.S. EPA's fish tissue criterion of 0.3 (.ig MeHg/g fish (0.3 ppm),
set to protect human health. The 2008 ISA reported on the negative impacts of Hg on the
development, morphology, survival, or reproduction in the following fish species:
walleye (Stizostedion vitreum), grayling (Thymallus thymallus), mummichog (Fundulus
heteroclitus), rainbow trout (Oncorhynchus mykiss), fathead minnows (Pimephales
promelas), and zebrafish (Danio rerio). However, a recent report on Hg in streams of the
U.S. by the USGS summarizes current research indicating that birds, fish, and fish-eating
wildlife experience negative effects of Hg at lower concentrations than the 0.3 ppm
criterion set to protect human health on the basis of fish consumption.
The 2008 ISA documented a link between decreased S deposition and decreased fish
MeHg concentrations. Recent research in Voyageurs National Park (a Class I Area)
supports this finding. There is also supporting evidence from fish surveys of Texas
reservoirs across regions with different S deposition loads, and from an S addition
experiment in a peat bog in the Marcell Experimental Forest in northern Minnesota,
where increased S loading increased Hg concentrations in larval Culex spp. (mosquitoes),
June 2018
94
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
which are an important food source for both aquatic and terrestrial species
(Appendix 12.4 and Figure 12-18). In addition to the studies that consider S deposition,
there are recent studies that consider sulfate concentrations in water in relation to fish Hg
concentrations in six lakes in South Dakota, and in the marshes of the Everglades
(Appendix 12.4). In the freshwater marshes of the Everglades, recent work indicates a
concentration of 1 mg/L sulfate to keep water MeHg low (Appendix 12.3.4.3). which is
also a sulfate concentration that will protect fish from elevated Hg burdens in that system
(Figure 12-14).
1.9.3 National-Scale Sensitivity and Critical Loads
The 2008 ISA identified ecosystems in the Northeast as particularly sensitive to Hg
methylation in response to S deposition, as many watersheds in this region have abundant
wetlands and freshwater water bodies with high DOC and low pH. The U.S. EPA
national stream surveys found that MeHg in predator fish exceeded the Hg criterion in a
quarter of stream miles and half the lakes surveyed. Fish MeHg levels were highest in
streams in watersheds with considerable wetland area, and surveys showed highest fish
MeHg concentrations in the southeastern U.S., suggesting that ecosystems sensitive to
SOxdeposition effects upon Hg methylation extend beyond the Northeast (Figure 12-15).
Recent studies confirm that Hg methylation is more widespread than was documented at
the time of the 2008 ISA. New research conducted in agricultural wetlands in California
suggests Hg methylation in these systems may provide a route to animal and human Hg
exposure through food, specifically MeHg concentrations in rice seeds.
There are no CLs for S to prevent sulfide phytotoxicity or Hg methylation, although there
are sulfate and sulfide water quality values that represent protective levels against toxic
effects of sulfide and Hg to biota (see Table 12-12). There are European CLs for Hg
concentrations in soil and fish tissue targeted to protect human health, drinking water
quality, and terrestrial soils, but these CLs are not framed in terms of SOx, Hg, or PM
deposition (see Appendix 12.6).
1.10 Ecological Effects of Particulate Matter other than N and S
Deposition
Since publication of the 2009 PM ISA, new literature builds upon the existing knowledge
of ecological effects associated with PM components other than those associated with N
and S deposition, especially metals and organics. In some instances, new techniques have
enabled further characterization of the mechanisms of PM on soil processes, vegetation,
June 2018
95
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
and effects on fauna. New studies provide additional evidence for community-level
responses to PM deposition, especially in soil microbial communities. However,
uncertainties remain due to the difficulty in quantifying relationships between ambient
concentrations of PM and ecosystem response. Overall, the body of evidence is
sufficient to infer a likely causal relationship between deposition of PM and a
variety of effects on individual organisms and ecosystems, based on information from
the previous review and new findings in this review.
PM deposition comprises a heterogeneous mixture of particles differing in origin, size,
and chemical composition. Exposure to a given concentration of PM may, depending on
the mix of deposited particles, lead to a variety of toxic responses and ecosystem effects.
Effects of PM on ecological receptors can be both chemical and physical (U.S. EPA.
2009a. 2004). As described in the 2009 Integrated Science Assessment for Particulate
Matter (2009 PM ISA), particulates that elicit direct and indirect effects on ecological
receptors vary in terms of size, origin, and chemical composition. Particle composition is
attributed to ecological outcomes to a greater extent than particle size (Grantz ct al..
2003). PM associated metals and organics are linked to responses in biota; however, the
heterogeneous nature of PM composition and distribution coupled with variability
inherent in natural environments confound assessment of the ecological effects of
particulates. Although most effects are from chemical composition of PM, there are some
effects of particle size generally limited to flux of solar radiation and soiling of leaves by
large coarse particles in areas near industrial facilities.
In general, new studies on PM deposition to vegetation support findings in previous PM
reviews on altered photosynthesis, transpiration, and reduced growth. Additional
characterization of PM effects at the leaf surface since the 2009 PM ISA has led to a
greater understanding of PM foliar uptake. Alterations in leaf fatty acid composition are
associated with metals transferred to plant tissues from PM deposition on foliar surfaces
(Appendix 15.4.2).
An important characteristic of fine particles (0.1 to 1.0 |im) is their ability to affect the
flux of solar radiation passing through the atmosphere, which can be considered in both
its direct and diffuse components. A newly available research method links changes in
expression of proteins involved in photosynthesis to changes in radiation associated with
aerosols and PM. Although this method has not been widely applied, it may represent an
important way to study mechanistic changes to photosynthesis in response to more
diffuse radiation resulting from PM in the air column (Appendix 15.2).
Several studies published since the 2009 PM ISA show PM chemical constituent effects
on soil physical properties and nutrient cycling. Previous findings in the PM ISA of
changes to microbial respiration and biomass are further supported by new studies.
96
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
Microbial community responses to PM vary in tolerance to heavy metals and organics
(Appendix 15.5.3V
In fauna, results from ecotoxicity assays with PM extracts using bacteria, rotifers,
nematodes, zebrafish, and earthworms support findings in the 2009 PM ISA that toxicity
is not related to the total mass of PM in the extract, but to the chemical components of the
PM. In nematodes exposed to PM from air filters, the insulin-signaling pathway was
identified as a possible molecular target. Use of wildlife as PM biomonitors has been
expanded to new taxa since the last PM review. Several studies in invertebrates and birds
report physiological responses to air pollutants, including PM (Appendix 15.6).
For ecosystem level effects, a gradient of response with increasing distance from PM
source was reported in the 2009 PM ISA. Newly available studies from long-term
ecological monitoring sites provide limited evidence for recovery in areas such as those
around former smelters due to the continued presence of metals in soils after operations
ceased. A novel experimental microecosystem using microbial communities living in
terrestrial mosses indicates that PM deposition alters responses of primary producers,
decomposers, and predators (Appendix 15.3).
1.11 Recovery of Ecosystems from Nitrogen (N) and Sulfur (S)
Deposition in the U.S.
Evidence from across the U.S. of ecosystem recovery from N nutrient enrichment and
acidification corresponding to long-term trends in N and S emissions is variable. Most
studies of recovery focus on ecosystem acidification recovery due to decreases in S
emissions and deposition. Overall N emissions and deposition have been increasing or
relatively steady; consequently, there has been little reported on N enrichment recovery
1.11.1 Overarching Concepts of Ecological Recovery from Acidification
Both chemical and biological indicators are used to assess the degree of ecological
degradation associated with environmental stressors and document responses in
ecosystems where improved conditions allow for recovery. Recovery can be documented
by measurement of indicators and projected/modeled recovery trajectories.
For acidification caused by N and S deposition, chemical recovery of aquatic and
terrestrial ecosystems are characterized by trends in water quality indicators (NOs .
SO42 , pH, ANC, inorganic monomeric Al, MeHg) towards inferred preindustrial values
or, in the case of inorganic Al and MeHg, below water quality threshold values protective
June 2018
97
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
of biota and human health. Preindustrial conditions varied across the U.S. in response to
variation in climate, geology, and biological communities, and preindustrial chemical
indicator values are currently inferred from models, paleolimnology samples, or historical
samples. When evaluating ecosystem recovery to acidification, it is important to note that
different chemical pools within the soil may recover at different rates with the same
decreases in declining atmospheric deposition at different rates. For example, soil
solution Ca: A1 ratio or SO42 or NO3 respond more quickly than will total N. Indicators
of slowly recovering pools (such as the percentage of base saturation in the soil or soil C
to N ratio) will have long response times with regard to changes in atmospheric
deposition. An indicator such as acid-neutralizing capacity (ANC), which is influenced
by both fast (solution) and slow (soil) pools, has an intermediate response time. Chemical
indicators such as ANC or pH may not necessarily follow a recovery path that mirrors the
reverse of the acidification path due to dynamic relationships among ANC, pH, DOC,
and inorganic Al; depletion of soil base cation pools; and/or pH-dependent S adsorption
on soils. In addition, the ANC level that reflects recovery of pH or Al may differ between
the acidification and recovery phases (Hcsthagcn et al.. 2008).
Biological recovery may follow chemical recovery of such water and soil quality
constituents; however, there may be a lag of decades between the onset of chemical
recovery and biological recovery KU-S. EPA. 2008a); Appendix 81. As observed in some
of the early studies on formerly acidified systems, the biological recovery trajectory may
exhibit hysteresis, where a system does not follow the same path from acidification to
recovery (Frost et al.. 2006). Complete biological recovery would entail a return to the
same species make-up, richness, and abundance as existed in the ecosystem in question
prior to the advent of human-caused acidic deposition (around the Year 1860 in North
American ecosystems). In a practical sense, complete biological recovery is probably not
attainable at most acidified locations within a reasonable management time frame
(perhaps 100 years) because soil reserves of base cations at many locations have been
depleted in response to many decades of acidic deposition and because other stressors, in
addition to acidic deposition, have also altered ecosystem structure and/or function or
will do so in the coming decades. Such stressors include changes in climate, land use, and
other perturbations. More commonly, only partial biological recovery may be possible.
Ecosystems deemed to be on a recovery trajectory are those found to be moving towards
a mix of species presence and abundance that approximates the undisturbed state. There
is substantial evidence that recovery rates from acidification differ between taxonomic
groups (e.g., rotifers vs. crustaceans) (Frost et al.. 2006; Mallev and Chang. 1994). In
general, recovery in freshwater ecosystems is characterized by populations of plankton
and benthic invertebrates prior to the recovery of fish populations, although most
biological communities studied to date have not returned to preacidification conditions,
even after recovery of chemical parameters.
98
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
1.11.2 Acidification Recovery in the U.S.
Long-term monitoring has been very important in tracking the ecological response to N
and acidifying deposition (Appendix 7 and Appendix 4.4). Experimental liming studies
have also provided some evidence for biological recovery, although these types of studies
are limited in the U.S. (Appendix 4.3.4 and Appendix 8.4.6). The historical focus on
aquatic acidification has resulted in more data to evaluate recovery in aquatic than
terrestrial ecosystems (Appendix 7.1.3). Fewer studies have tracked the potential
recovery of terrestrial ecosystems; however, since the early 1990s, evidence has grown,
which indicates that soils in some areas are beginning to recover, yet most sensitive
regions continue to acidify in response to deposition (Appendix 4.6.1). In areas where N
and S deposition has decreased, chemical recovery must first create physical and
chemical conditions favorable for growth, survival, and reproduction in order for
biological recovery to occur.
The northeastern U.S. and southern Appalachians are two regions of the U.S. where a
large body of research has evaluated recovery. In the Northeast, evidence for chemical
recovery is primarily from soils (Appendix 4.6.1) and freshwater lakes and streams
(Appendix 7.1.5.1). In regards to biological recovery (Appendix 8.4). newer studies have
documented some evidence for zooplankton recovery and the successful reintroduction of
brook trout in previously acidified Adirondack water bodies or recolonization of
previously acidic lakes from refugia (Appendix 8.6.6). In addition to decreased
acidification, a few studies report declines in methylmercury concentrations in biota or
water in response to decreasing S, which is suggestive of ecosystem recovery
(Appendix 12.5).
In contrast to the northeastern U.S., there is little evidence for recovery in the southern
Appalachian Mountain region (Appendix 4.6.1 and Appendix 16.3). This area is
characterized by an abundance of low-ANC streams situated on acidic, highly weathered
soils. Streams in this region are strongly affected by SOr adsorption on soils, and
long-term monitoring studies suggest that soil base cation depletion has prevented
chemical recovery (Appendix 7.1.5.1.4). Biogeochemistry modeling scenarios suggest
that even with large decreases in SO42 deposition, it may take decades for soil base
cation levels to recover in this region.
New studies continue to support findings in the 2008 ISA that biological response to
water chemistry recovery varies among taxa and among water bodies, and that most
biological communities studied have not returned to preacidification conditions, even
after recovery of chemical parameters (Appendix 8.4). Since the 2008 ISA, research has
described the DOC of many lakes and streams has risen, with the source of the DOC
likely from the soils in the adjacent terrestrial watershed (Chapter 1, Table 1-2;
June 2018
99
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Appendix 4.3.9 and Appendix 7.1.2.9). The mechanism causing the observed increase in
DOC is unclear, and may be due to a combination of soil recovery from acidification,
changes in climate (e.g., temperature and precipitation), and N deposition, among other
mechanisms. DOC interacts like a weak acid; therefore, DOC content may affect pH and
ANC levels and constrain the extent of recovery from acidification. At the same time, the
acidic properties of DOC make it a host for binding trace metals such as toxic inorganic
A1 (for additional discussion on inorganic A1 and DOM see Appendix 4.3.5) and
decreases the toxicity of dissolved A1 to aquatic organisms. Overall, current research
indicates DOC increases are inconsistent across surface waters in the U.S. with large
increases in DOC with acidification recovery in some locations while other recovering
sites have not shown increasing trends in DOC.
1.11.3 Nitrogen (N) Driven Nutrient Enrichment Recovery in the U.S.
In most freshwater systems sensitive to nutrient effects of atmospheric deposition of N,
there is no evidence for biological recovery although decreases in NOs concentrations
consistent with declines in N deposition have been reported in some regions of the U.S.,
notably the Appalachian, Adirondack, and Rocky Mountains (Appendix 7.1.5). Some
estuaries have shown improvements in biological indicators, such as increases in the
extent of SAV, in response to decreases in N inputs from atmospheric deposition and in
wastewater and agricultural runoff. For an example, see the Tampa Bay case study
(Appendix 16). In other coastal areas of the U.S., biological indicators of nutrient
enrichment have remained relatively unchanged or declined. In the well-studied
Chesapeake Bay watershed where extensive restoration efforts have been implemented,
water quality and measures of ecological condition have shown little improvement during
a 23-year period (Williams et al.. 2010). The one exception to the pattern of no
improvement in water quality was an observed increase in the amount of SAV
(Appendix 10.2.5).
1.12 Climate Modification of Ecosystem Response to Nitrogen (N)
and Sulfur (S) Addition
Nitrogen and S loading occurs in many ecosystems concurrently experiencing multiple
stressors, including human-driven climate change. Climate change effects on U.S.
ecosystems were recently summarized in the U.S. National Climate Assessment
(Galloway et al.. 2014; Groffman et al.. 2014). Each appendix of the ISA evaluating N
enrichment or acidification includes a section on how climate modifies the ecosystem
response. Additionally, to serve as a foundation for the discussion, text in Appendix 13 is
June 2018
100
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
excerpted from Greaver et al. (2016). a current review of how climate (e.g., temperature
and precipitation) modifies ecosystem response to N that focuses on empirical
observations.
Anthropogenic emissions of greenhouse gases are likely to cause a global-average
temperature increase of 1.5 to 4.0°C and a significant shift in the amount and distribution
of precipitation by the end of the 21st century (Collins et al.. 2014). Recent work has
focused on the effects of anthropogenic N on the Earth's radiative forcing (Pinder etal..
2012) and how temperature and precipitation alter ecological responses to N exposure
(Greaver et al.. 2016). Most work is conducted on the effects of climate and N or
acidifying deposition (N + S); relatively little work is conducted on how climate modifies
ecosystem response to S.
Climate effects on ecosystems is a rapidly expanding field; however, for some processes,
we are beginning to understand how temperature and precipitation may interact to alter
ecosystem response to N and S addition. Although, for many biogeochemical pools and
processes data is insufficient to quantify either the direction or magnitude of how climate
may alter ecosystem response with certainty.
1.13 Key Scientific Uncertainties
Evaluation of uncertainty is an important part of ecosystem assessment. Uncertainty
refers to the absence of information and is a way to describe how certain we are in
scientific knowledge. As described by Curry and Webster (2011). the nature of
uncertainty can be expressed by the distinction between ontic uncertainty and epistemic
uncertainty. Ontic uncertainty is associated with inherent variability or randomness, and
is an irreducible form of uncertainty. Epistemic uncertainty is associated with
imperfections of knowledge, which may be reduced by further research and empirical
investigation. Walker et al. (2003) [as summarized in Curry and Webster (2011)1
characterized uncertainty as a progression from deterministic understanding to total
ignorance:
"Statistical uncertainty is the aspect of uncertainty that is described in
statistical terms. An example of statistical uncertainty is measurement
uncertainty, which can be due to sampling error or inaccuracy or
imprecision in measurements.
Scenario uncertainty implies that it is not possible to formulate the
probability of occurrence of one particular outcome. A scenario is a
plausible but unverifiable description of how the system and/or its
June 2018
101
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
driving forces may develop over time. Scenarios may be regarded as a
range of discrete possibilities with no a priori allocation of likelihood.
Recognized ignorance refers to fundamental uncertainty in the
mechanisms being studied and a weak scientific basis for developing
scenarios. Reducible ignorance may be resolved by conducting further
research, whereas irreducible ignorance implies that research cannot
improve knowledge."
The understanding and reporting of uncertainty is not consistent across scientific
disciplines, and uncertainty may be quantified by various methods. Csavina et al. (2017)
provided an overview of terminology and definitions of 41 different terms used to
describe uncertainty. Here we provide a brief summary of some of the key methods that
may be used to evaluate the uncertainty of the relationship between NOx, SOx, and PM
pollutants and ecological effects. Quantified estimates of uncertainty vary according and
number of decision points (Chapter 1.13.2.3) including the method used and the input
parameters under consideration; therefore, the analyses and discussion of quantified
uncertainty values will occur in the Risk and Exposure Assessment as scoped in the 2017
IRP (U.S. EPA. 2017b).
1.13.1 Atmospheric Science
Estimating atmospheric deposition involves quantification of emissions, atmospheric
concentrations, and deposition fluxes of the various species that make up atmospheric
SOx, NOy, and NHx. This is accomplished with environmental measurements, model
predictions, or hybrid approaches that combine measurements and modeling methods.
There are a wide range of uncertainties across the environmental measurements and
model parameters used to estimate atmospheric deposition fluxes. The largest
uncertainties are those for dry deposition and ammonia emissions, whether measured or
modeled. The smallest uncertainties are associated with ambient concentration
measurements and continuously monitored stationary emissions like electric power
plants.
1.13.1.1 Emissions Uncertainty
Quantitative uncertainty estimates are not documented in the NEI, but uncertainties are
often evaluated through separate efforts by comparing inventory predictions with
measured long-term trends, statistical source apportionment methods, inverse chemical
transport modeling, and comparison with satellite data (Appendix 2.2.2). SO2 and NOx
June 2018
102
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
emission uncertainties for electric power generating units, the major source of SO2 and an
important source of NOx, are in the 10-15% range because emissions are usually
continuously monitored (Appendix 2.2.3). NOx emission uncertainties for mobile
sources, the largest source of NOx, arise from differences in engine type, size, age, and
maintenance, as well as fuel composition and emission control equipment.
Over-estimation of NOx emissions from mobile sources was proposed as an explanation
for modeled NOx concentration bias in several studies. However, mixed results have
been observed across several studies when modeled concentrations were compared to
measurements. Estimates of NOx emissions uncertainties are in the 10-20% range for
on-road gasoline and diesel vehicles, and up to 30% for off-road vehicles like ships,
airplanes, and locomotives (Appendix 2.2.3). In contrast, total NH3 emissions
uncertainties appear to be greater, underestimated by as much as a factor of two or more
according to several recent studies (Appendix 2.2.3). The predominant sources, livestock
operations and fertilizer application, exhibit large temporal and regional variability due to
differences in climate conditions and farming practices. As a result, detailed models are
required for estimating NH3 emissions (Appendix 2.2.2). but data on local environmental
conditions and farming practices necessary for good model performance are often not
available. Large discrepancies between modeled and measured N concentrations and
deposition rates have been attributed to uncertainties in ammonia emissions
(Appendix 2.2.3).
1.13.1.2 Atmospheric Measurement Uncertainty
Uncertainties in concentration and deposition measurements from network-based
measurements are generally under 20%, and surface concentration uncertainties from
satellite-based measurements typically somewhat higher. Concentration and deposition
data are derived from several specialized national monitoring networks, including the
national SO2 monitoring network, the NCore network for multipollutant concentration
monitoring including NOy, the Ammonia Monitoring Network, CASTNet for estimating
dry deposition, and the National Trends Network for wet deposition (Appendix 2.4.1).
Uncertainties are estimated from reports of precision in data quality reports where
available, and otherwise from network data quality objectives.
For air concentration measurements used for estimating dry deposition, CASTNet,
measured precision was 2-5% for sulfate, 5-13% for nitrate, and 2-6% for ammonium in
2016 (Appendix 2.4.5). Additional uncertainty is associated with for estimating dry
deposition from NTN concentration data. Uncertainties of 30% for SO2 and 40% for
HNO3 have been reported using a simple inferential approach (Clarke et al.. 1997).
However, single site determinations are of limited use because of dry deposition fluxes
June 2018
103
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
are determined by a number of factors and can vary considerably over small spatial
scales. In most recent efforts, dry and total deposition on a regional or national scale is
usually modeled with CTMs (Chapter 1.13.1.3).
Precipitation concentration measurement precision and estimated wet deposition
precision in the National Trends Network were less than 7% for sulfate and nitrate and
less than 20% for ammonium. Ammonia air concentration measurement methods used in
AMoN evaluated a precision of 10% (Appendix 2.4.6). Minimum performance
specifications for SO2 monitoring from the national SO2 monitoring network include a
precision of 2.0% (Appendix 2.4.4). Data quality objectives for NOy in the NCore
network include a precision of 15% (Appendix 2.4.2). Uncertainty in satellite-based
measurements depend on vertical profile, cloud fraction, cloud pressure, surface
reflectivity, and extent of aerosol scattering. Estimates of 20% for NO2 (Appendix 2.4.2)
and 10-45% for SO2 (Appendix 2.4.4) have been reported for cloud free conditions.
1.13.1.3 Atmospheric Modeling Uncertainty
The Community Multiscale Air Quality modeling system is probably the most widely
used model in the U.S. for estimating atmospheric deposition. In a recent CMAQ
evaluation, atmospheric nitrate concentrations were overestimated by 22 to 26%, sulfur
dioxide concentrations were overestimated by 39 to 47%, and sulfate concentrations were
underestimated by 9 to 17%, as annual averages over a range of 4 years compared to
surface-based measurements (Appendix 2.5.3). Mixed results have been observed in
several recent comparisons of CMAQ wet deposition estimates to network-based
measurements, with average differences in modeled results and measurements ranging
from 15% or less to 99% for nitrate, sulfate, and 15% or less to 60 % for ammonium
(Appendix 2.5.3). Modeling methods for estimating dry and total deposition are still
under development, and uncertainties have not been extensively evaluated or quantified.
Recent sensitivity analysis results found less than 5% differences in total deposition
estimates because of compensation of competing model processes, but extensive
comparison of model results and measurements are not available (Appendix 2.5.3).
In addition to measurable uncertainties associated with measurement precision or
comparisons between models and measurements, there are also structural uncertainties
due to incomplete understanding of the underlying science related to atmospheric
deposition that are not possible to quantify. The main structural uncertainties associated
with deposition estimates are canopy effects of NOx (including both bidirectional gas
exchange and canopy reactions), bidirectional exchange of NH3 with biota and soils, and
June 2018
104
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
processes determining transference ratios that relate average concentration to deposition
(Appendix 2.5).
1.13.2 Ecological Effects
Evaluation of ecological effects caused by acidification or eutrophication involves a suite
of parameters and dose-response functions, both empirical and modeled. The quantitative
uncertainty of empirically observed variables in ecology is determined by the use of
statistics. A suite of mathematical statistical models are available to describe the
variability among empirical observations and the strength of a cause and ecological effect
relationship, the appropriate method to apply depends on the experimental design.
Statistics for empirical data include calculation of probability, distributions, standard
deviation, variance, /-tests. ANOVA, linear regression, spatial statistics, Bayesian
analysis, and multivariate analysis, among others. In general, ecological endpoints
determined by empirical studies to be affected by deposition were reported in the ISA if
they were statistically significant; this means the magnitude of effect was larger than the
estimated uncertainty.
Models of chemical and ecological processes, including biogeochemistry, provide
representations of biological and geochemical interactions through mathematical
expressions. The models used to characterize aquatic and terrestrial biogeochemistry
response to N and S deposition can be complex, including many interacting variables.
Model results are often compared to empirically collected data to confirm the model.
Each of the input variables used in a biogeochemical model entails uncertainty. Model
uncertainty is governed, in part, by how close the model predictions are to actual
observations. Uncertainty in modeled results may arise from limitations in input data or
from limitations in model assumptions. Statistical inference methodologies enable
uncertainty analysis and determine the strength of the relation between a given uncertain
input and the output (i.e., sensitivity analysis). For biogeochemistry models these
methods include first-order sensitivity index, Monte Carlo technique, extended Fourier
amplitude sensitivity test, Morris one-factor-at-a-time, and Bayesian analysis.
1.13.2.1 Empirical Critical Loads
Empirical N CLs for terrestrial and aquatic ecosystems reported in this ISA have been
estimated using empirical data sets. The exact effects threshold may be determined using
expert judgement. For example, if three levels of N addition are applied to a study site
(10, 20, and 30 kg N/ha/yr) and an effect is noticed at 20 kg N/ha/yr, then the CL is
June 2018
105
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
estimated at <20 kg N/ha/yr. Another approach would be to fit a mathematical function to
the observations, and a scientific judgement made to identify the level of deposition
and/or N addition, or threshold, at which the ecological effect is considered to occur and
which is likely to be biologically adverse.
There are some challenges associated with developing CLs that can result in uncertainty.
First, because biological responses are often continuous, there can be a lack of an obvious
cut off between adverse and nonadverse effects. As a result, individual author groups
have selected different response thresholds. For example, N CLs for lichens have been
calculated for (1) deposition values associated with thallus N concentrations above the
97% distribution quantile observed for clean sites (Fcnn et al.. 2008). (2) community
composition shifts from oligotroph to eutroph dominance (Tenn et al.. 2008). (3) low
probability of detecting regionally distributed sensitive species (Root et al.. 2015; Geiser
et al.. 2010). or (4) extirpation of oligotrophs (Fcnn et al.. 2008). Secondly, clean site data
can be lacking in some ecoregions. For instance, few empirical data are available for sites
in the eastern U.S. with deposition rates <4 kg N/ha/yr. This makes it difficult to quantify
physiological or community compositional that may have occurred in this region at
deposition rates of 1-4 kg N/ha/yr.
The Pardo et al. (2011a) study provided a compilation of terrestrial and aquatic N CLs
reported since the 2008 ISA. Uncertainty in derivation of empirical CLs for N input as
presented by Pardo etal. (2011a) arises in estimation of the ambient (and perhaps
historical) deposition loads and in estimation of the biological effects caused by those
deposition levels. According to Pardo et al. (2011a). sources of uncertainty in N
deposition estimates for N CLs at the Ecoregion Level 1 scale include: "(1) the difficulty
of quantifying dry deposition of nitrogenous gases and particles to complex surfaces;
(2)	sparse data, particularly for arid, highly heterogeneous terrain (e.g., mountains); and
(3)	sites with high snowfall or high cloud water/fog deposition, where N deposition tends
to be underestimated." Examples of high uncertainty include high-elevation sites in the
Rockies and Sierra Nevada mountains, due in part to highly uncertain estimates of dry
deposition (Appendix 2). For sensitive receptors such as phytoplankton shifts in high
altitude lakes, N deposition model bias may be close to, or exceed, predicted CL values
(Williams et al.. 2017a).
Physical, chemical, and ecological variability across lakes affect their response to N
deposition and contribute to uncertainty of CL estimates (Appendix 9.1.1.2). A review by
Bowman et al. (2014) noted that current N CLs for sensitive alpine systems may not be
protective under future climate scenarios of warmer summer temperatures and a shorter
duration of snow cover.
106
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Between the publication of Pardo etal. (2011a) and the cutoff date for literature in this
ISA (May 2017), some additional aquatic and terrestrial N CLs have been published
(Appendix 4; Appendix 6.5). Simkin et al. (2016) was not based on field addition or N
gradient of deposition studies, instead the methods were a spatial analysis of plant
diversity using a large data set of over 15,000 forest, shrubland, and herbaceous sites
across the U.S. Atmospheric N deposition varied nearly 20-fold across the site gradient.
They found that N deposition was negatively correlated with plant species richness at
many locations, but positively correlated at others. In one large empirical data set used to
describe N CLs for U.S. grassland biodiversity, Simkin et al. (2016) estimated the
uncertainty surrounding the mean CL estimates. For open canopy ecosystems, for
example, they estimated a mean of 8.7 kg N ha/yr and provided 95% confidence
intervals, which can be used as estimates of uncertainty, of 6.4 to 11.3 kg N ha/yr. For
closed canopy systems, the mean of 13.4 kg N ha/yr was surrounded by a 95%
confidence interval of 6.8 to 22.2 kg N ha/yr.
The majority of studies that evaluate terrestrial N CLs for N enrichment effects are based
on observed response of a biological receptor to N deposition (or N addition as a proxy
for deposition), without a known soil chemistry threshold that causes the biological
effect. In contrast, critical loads for acidification are typically based on the deposition
amount that causes a value of a soil chemical indicator which is known to cause adverse
biological effect. The link between soil chemical indicator and biological effect is based
on empirical evidence (Appendix 5). The relationship between deposition and the
biogeochemistry that causes effects on soil chemistry is typically modeled (Appendix 4.
Chapter 1.13.2).
1.13.2.2
1.13.2.2.1
Modeled Critical Loads
Terrestrial and Aquatic Acidification: Biogeochemistry
A variety of process models have been used to estimate past and future resource
conditions under scenarios of acidification/recovery responses and critical and target
loads, both aquatic and terrestrial. Models include simple approaches such as the simple
mass-balance equation (SMBE), and dynamic models, such as PnET-BGC and ForSAFE,
MAGIC, VSD, and VSD+ (Appendix 4.5). Critical loads for terrestrial and aquatic
acidification are calculated by the model to determine the amount of deposition that alters
soil or water chemistry to a threshold value known to have detrimental effects on a
biological receptor.
June 2018
107
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Of the several well-established models of terrestrial biogeochemistry used to evaluate soil
acidification (Appendix 4.5). each relies heavily on input or simulated values for base
cation weathering (BCw) rate, one of the most influential yet difficult to estimate
parameters in the calculation of critical acid loads of N and S deposition for protection
against terrestrial acidification (Appendix 4.5.1.1). Obtaining accurate estimates of
weathering rates is difficult because weathering is a process that occurs over very long
periods of time, and estimates on an ecosystem's ability to buffer acid deposition rely on
accurate estimates of weathering. Various approaches can be used to estimate BCw,
including the empirical soil clay approach, the PROFILE model (e.g.. Phclan ct al..
2014). the F-factor approach (U.S. EPA. 2009c). and calibration of a dynamic model such
as MAGIC (e.g., McDonnell et al.. 2014b; Povak et al.. 2014). There are new studies on
estimating BCw, including evaluation of uncertainty (Whitfield. 2018; Futter ct al..
2012). When applying PROFILE to upland forests in the U.S., Whitfield (2018) found
the greatest uncertainty in BCw estimate was due to the particle size class-based method
used to estimate the total specific surface area upon which weathering reactions can take
place.
The uncertainty of forest soil CLs for acidification in U.S. calculated using simple
mass-balance equations (SMBE) was investigated by Li and McNultv (2007). The results
included a quantification of how the 17 of the model's parameters contributed to the
uncertainty and indicated that uncertainty in the CLs came primarily from components of
base cation weathering and acid-neutralizing capacity, whereas the most critical
parameters were BCw, base rate, soil depth, and soil temperature. The authors concluded
that improvements in estimates of these factors are crucial to reducing uncertainty and
successfully scaling up SMBE for national assessments (see Appendix 4.6).
Several dynamic models are commonly used to model terrestrial soil acidification
(Appendix 4.5). Tominaga et al. (2009) conducted a Monte Carlo multiple-model
evaluation of the dynamic models MAGIC, SAFE, and VSD and found that given the
same deposition scenario, the three models (without calibration) simulate changes in soil
and soil solution chemistry differently, but the basic patterns were similar. The authors
also found the greatest differences in model outputs were attributed to the cation
exchange submodel. Bonten et al. (2015) compared how well the common types of
dynamic models used to evaluate terrestrial soils (VSD, MAGIC, ForSAFE, and
SMARTml). quantified several variables including soil S, soil pH, soil ANC, BC, base
saturation, and Al (Appendix 4.5.3).
Uncertainty analysis of a dynamic model (VSD) used for CL based on soil chemistry
chemical limits showed that the main drivers of uncertainty were largely dependent on
the chemical criterion selected [Appendix 5.5.3.3; (Reinds and de Vries. 2010)1. For
108
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
example, base cation weathering, deposition, and the parameters describing the H-Al
equilibrium in the soil solution were the main sources of uncertainty in the estimates of
maximum critical loads for S (CLmax[S]) based on the Al:Bc criterion of 1.0, and
uncertainty in CLmax(S) based on ANC was completely determined by base cation
inputs. The denitrification fraction was the most important source of uncertainty for the
maximum critical loads of N (CLmax[N]). Calibration of VSD reduced the levels of
uncertainty for all critical loads and criteria.
Fakhraei et al. (2017b) reviewed sensitivity and uncertainty analysis techniques
(e.g., first-order sensitivity index, Monte Carlo technique, extended Fourier amplitude
sensitivity test, Morris one-factor-at-a-time, and Bayesian analysis) in the context of a
biogeochemistry model. The authors apply these techniques to determine the uncertainty
and sensitivity of the PnET-BGC model calculation of TMDLs of acidifying deposition
that occur in high-elevation, acid-impaired streams in GSMNP (Fakhraei et al.. 2017a).
Sensitivity analyses showed that modeled estimates of maximum allowable acidifying
deposition loads were most sensitive to uncertainty in model input parameters of air
temperature, precipitation quantity, and rate of calcium weathering. Importantly, as more
uncertainty was incorporated into model input parameters (±5 to ±10 to ±20%
uncertainty), estimates of allowable deposition loads to protect aquatic ecosystem
recovery decreased in magnitude (Fakhraei et al.. 2017a).
Biogeochemistry and Plant Biodiversity Linked Modeling
Plant biodiversity models, such as VEG and PROPS, have been coupled to dynamic
biogeochemical models, such as ForSAFE and VSD+ (Mcdonnell et al.. 2018b;
Mcdonnell et al.. 2018a; Phelan et al.. 2016). ForSAFE-Veg is an older and more broadly
applied model than VSD + PROPS. There are some key differences between Veg and
PROPS. Plant species in the VEG component of ForSAFE-VEG are defined by
mathematical equations based on expert opinion regarding such parameters as plant needs
for moisture, sunlight, and N supply to represent unobservable fundamental niches. In the
PROPS statistical relationships based on empirical data are used to characterize plant
species, which is likely more approximate of real world niches influenced by competition
among species. These model chains are subject to the same constraints and uncertainties
as the biogeochemical models on their own, plus those of the plant response modules.
Aquatic Eutrophication Modeling
Many of the models that estimate N loads to the coastal zone from the landscape and
freshwater inflow have been compared, and there is a good deal of knowledge about their
109
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
limitations and uncertainties (McCrackin et al.. 2013; Alexander et al.. 2008). In a 2000
National Research Council review, it was determined that these models are
hydrodynamically complex and tend to be specific to particular sites. Thus, they are
difficult to apply broadly (NRC. 2000).
This SPARROW model application used only wet deposition as the measure of N
deposition. A large amount of N from nonpoint source urban influences (this is most
likely primarily the dry deposition of exhaust N gases) often approximately doubles the
importance of N deposition as an N source to higher order river systems (Howarth.
2008a. b).
1.13.2.3 Key Considerations
The choice of model for CL estimation, or for scenario projection, depends largely on the
availability of time, data, and resources. Major decisions inherent in the modeling efforts
include:
•	Empirical observation or application of a model
•	Steady state or dynamic model
•	Statistical or process-based model
•	Protection against acidification or nutrient N enrichment
•	Site-specific, regional, or national spatial scale
•	Resources to be protected (i.e., stream, lake, soil, vegetation, aquatic biota)
•	Chemical indicator(s) of adverse effects (e.g., water ANC, water NO3", soil BS)
•	Critical level(s) for selected indicator(s)
•	Time frame of evaluation (i.e., ambient, 2050, long-term steady state)
Each of these decision points introduces additional uncertainties, data needs, and
potential assessment errors. U.S. EPA (2008a) summarized CL research and monitoring
needs identified by U.S. EPA (2006b) at the time of the previous (2009) U.S. EPA Risk
and Exposure Assessment.
1.14 Ecosystem Services
For acidification, the ecosystem service literature since the 2008 ISA includes studies that
better characterize ecosystem service valuation by pairing biogeochemical modeling and
benefit transfer equations informed by willingness-to-pay surveys, especially for the
Adirondacks and Shenandoah regions (Appendix 14). Aside from valuation, the total
June 2018
110
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
number of ecosystem services affected by N and S deposition is better quantified by the
new studies that use the final ecosystem goods and services classification system
(FEG-CS). However, for many regions and specific services, poorly characterized
dose-response between deposition, ecological effect, and services are the greatest
challenge in developing specific data on the economic benefits of emission reductions
(NAPAP. 201IV
In the 2008 ISA there were no publications that specifically evaluated the effects of N
deposition on ecosystem services associated with N driven eutrophication. Since the 2008
ISA, several comprehensive studies have been published on the ecosystems services
related to N pollution in the U.S. (Appendix 14). These include an evaluation of services
affected by multiple N inputs (including N deposition) to the Chesapeake, a synthesis of
the cost-benefits on N loading across the nation, and analysis of the amount of N that
leaked out of its application causing effects on adjacent ecosystems and ecosystem
services. In their work (this work specifically identified the costs of the atmospheric
portion of total N loading), the estimate of the total number of ecosystem services
affected by N is better quantified by the new studies that use FEG-CS (Bell et al.. 2017;
Clark et al.. 2017; Irvine et al.. 2017; O'Dea et al.. 2017; Rhodes et al.. 2017). In these
analyses, critical load exceedances for N related air pollution were used as a model
stressor from which a total of 1,104 unique chains linking stressor to beneficiary were
identified.
The conclusions considering the full body of literature are that (1) there is evidence that
N and S emissions/deposition have a range of effects on U.S. ecosystem services and
their social value; (2) there are some economic studies that demonstrate such effects in
broad terms; however, it remains methodologically difficult to derive economic costs and
benefits associated with specific regulatory decisions/standards; and (3) thousands of
causal relationships are now documented between N and S air pollution and changes in
Final Ecosystem Goods and Services.
Ill
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
APPENDIX 1. INTRODUCTION TO APPENDICES
The appendices and Chapter 1 serve different purposes in this ISA. Chapter 1 is meant to
summarize the key messages derived from assessment of the policy relevant science of
NOy, SOx and PM in this review of the secondary National Ambient Air Quality
Standards. It provides a general introduction to the purpose, process, development, and
organization of the ISA as well as highlights connections, concepts, and changes based
on new evidence and causality. In addition, Chapter 1 provides a discussion of
uncertainty and a synthesis of information on the recovery of ecosystems from N and S
deposition.
While the purpose of Chapter 1 is to synthesize, integrate, and provide key messages, the
purpose of the appendices is to provide a more detailed description of the state of science
for specific topic areas. Appendix 1 is an introduction to the purpose and organization of
Appendix 2-Appendix 16. Appendix 2 characterizes the sources, atmospheric processes,
and the trends in ambient concentrations and deposition of NOy, SOx, and PM.
Appendix 3 describes the direct effects of NOy and SOx gases on plants and lichens.
Appendix 4-Appendix 6 describe the effects of N and S deposition on biogeochemistry
and the biological effects of acidification and N enrichment in terrestrial environments.
Appendix 7 describes the effects of N and S deposition on aquatic biogeochemistry.
Appendix 8-Appendix 10 characterize the biological effects of freshwater acidification,
freshwater N enrichment, and N enrichment in estuaries and near-coastal systems.
Appendix 11 describes the effects of N deposition on wetlands, and Appendix 12
characterizes the ecological effects of S as a nutrient. Appendix 13 presents information
on climate modification of ecosystem response to N and S, while Appendix 14 discusses
ecosystem services. Appendix 15 is a review of the ecological effects of forms of PM, not
related to N or S deposition. Finally, Appendix 16 presents case studies for six locations
in the U.S. (southern California, northeastern U.S., Rocky Mountain National Park,
southeastern Appalachia, Tampa Bay, and the Adirondacks) where data are sufficient to
well characterize the ecological effects of N and S deposition. These sites would
therefore make good candidates to better assess risk and exposure by exploring linkages
across various effects and ecosystems-types in a specific location.location.
June 2018
1-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
APPENDIX 2. SOURCE TO DEPOSITION
2.1. Introduction
In this appendix, emphasis is placed on those species subject to atmospheric processes
relevant for review of the air quality criteria and associated welfare-based secondary
National Ambient Air Quality Standards (NAAQS) for oxides of nitrogen, oxides of
sulfur, and/or particulate matter (PM). As such, this appendix largely focuses on
examining the fundamental and applied science of atmospheric processes relevant to
assessing environmental exposures and effects associated with atmospheric deposition of
nitrogen (N) and sulfur (S) species, including those present in PM.25 Together this
information serves as a prologue to the detailed descriptions of the evidence of ecological
effects from oxidized and reduced N and oxidized S, including those present in PM, that
follow in later appendices.
Recent advances in research on N and S emissions sources, atmospheric transformation
and transport, measurement and modeling techniques, atmospheric loadings, and
deposition processes relevant to this review of the NAAQS are evaluated in this
appendix. N and S species of interest are generally classified into three groups: oxidized
nitrogen, reduced nitrogen, and oxidized sulfur. While NO2 and SO2 are the most well
known as air pollutants, research on the entire range of oxides of nitrogen and sulfur is
considered for review of the air quality criteria. Reduced nitrogen is also discussed
because it strongly influences the atmospheric deposition of NO y and SOx as well as the
chemistry of PM formation. Oxidized nitrogen, reduced nitrogen, and oxidized sulfur all
have particulate forms (NO;, . NH4 . SO42 ). which together account for a large fraction
of PM mass (U.S. EPA. 2009a). as well as gas-phase components that act as major
precursors to PM. Thus, a consideration of the combined effects of oxides of nitrogen,
oxides of sulfur, and PM requires an understanding of the atmospheric processes
involving oxidized nitrogen, reduced nitrogen, and oxidized sulfur.
Oxidized nitrogen species considered here range from nitric oxide (NO) and nitrogen
dioxide (NO2), collectively referred to as NOx, to higher order organic and inorganic
oxidation products, collectively referred to as NOz (e.g., PNO3, HNO3, HONO, PAN,
other organic nitrates). NOz is especially relevant when considering nutrient addition to
ecosystems and the acidification of surface waters. NOx and NOz together are referred to
25 Since ecological effects of PM are governed mainly by PM composition, the most relevant PM species (nitrate
and sulfate) are also species that are derived from sulfur and nitrogen oxide precursors, and there is a high degree of
overlap in the discussion of the impacts of NOy and SOx, and the impacts of PM2 5. The PM ISA (U.S. EPA. 2009a)
provides more extensive information about the atmospheric processes for total PM mass.
June 2018
2-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
as NOy (i.e., NOx + NOz = NOy). Nitrous oxide (N2O) is an oxide of nitrogen, but it is
not included as a component of NOy. N2O contributes to stratospheric ozone depletion
and climate forcing [AR5; (IPC'C. 2013)|. but it contributes little to N deposition, and is
not included in this appendix.
Reduced nitrogen species are NH3 and NH4 as well as reduced organic nitrogen
compounds. NH3 and NH4 together are referred to as NHx (i.e., NH3 + NH4+ = NHx).
Reduced nitrogen contributes to acidification and N enrichment, and it also has a key role
in neutralizing acidity in cloud, fog, and rain water as well as aqueous aerosol particles
formed from atmospheric oxidation of SO2 and NOx. Additionally, NH3 is a precursor for
atmospheric particulate matter, reacting with gas-phase nitric acid (HNO3) to form
ammonium nitrate (NH4NO3), a major component of N deposition in many areas of the
U.S. For this assessment, NOy and NHx are grouped together as total reactive nitrogen,
Nr (i.e., NOy + NHx = Nr). Nr does not include nitrous oxide and reduced organic
nitrogen compounds. However, to the extent it is available, information on the sources,
abundances, and fate of reduced organic nitrogen is included in the sections that follow.
Gaseous oxides of sulfur (SOx) is defined to include sulfur monoxide (SO), sulfur
dioxide (SO2), sulfur trioxide (SO3), and disulfur monoxide (S2O). Of these only SO2 is
present in the lower troposphere at concentrations relevant for environmental
considerations. However, SO2 interacts with particles and cloud drops and is oxidized to
sulfate. SO2 and sulfate (SO42 ) account for much of the acidification of surface water in
the U.S. and together these make up total oxides of sulfur discussed in this appendix
(SOx).
Particulate matter (PM) impacts discussed in this document are also mainly focused on N
and S containing species, which together usually make up most of the fine PM mass in
many areas of the U.S. PM is usually classified into two size fractions which differ in
their physical and chemical characteristics, atmospheric behavior, and health and
environmental effects. These are PM2 5, particles smaller than 2.5 |_im in diameter, and
PM10-2 5, particles with diameter between 2.5 and 10 |_im. Ecological impacts of PM
depend largely on its composition (U.S. EPA. 2009a. 2004). Together PM2 5 and PM10-2.5
make up PM10. PM contains numerous individual components representing a wide range
of chemical and physical properties. However, in most areas of the U.S. PM2.5 mass is
composed almost entirely of sulfate, nitrate, and organic materials. In contrast, PM10-2.5 is
composed of crustal material similar in composition to soil in the area where the PM10-2.5
is found, as well as sea salt in coastal areas. There is little discussion of PM10-2.5 effects in
this document because in most rural and remote areas PM10-2.5 is largely due to natural
sources like soil and sea salt.
2-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
There are several other reasons for focusing on the sulfate and nitrate fraction of PM2 5
mass in this document in addition to the observation that they often account for most
PM2 5 mass. Together they also usually account for an even greater fraction of PM2 5 in
rural and remote areas that make up most of the U.S. land mass over which effects in this
document are relevant. While organic matter can also account for a large fraction of
PM2 5, it is composed of a wide variety of individual compounds that cannot be identified
at a molecular level, making it difficult to assess ecological impacts. Also, there is little
information on organic PM impacts, except for individual compounds that make minor
contributions to mass. As a result, the main contributors to PM2 5 mass for which
ecological impacts can be readily assessed are limited to sulfate and nitrate, which are
also components of total oxides of sulfur and oxides of nitrogen, respectively. Since
ecological effects of PM are governed mainly by PM composition, the most relevant PM
species (nitrate and sulfate) are also species that are derived from sulfur and nitrogen
oxide precursors, and there is a high degree of overlap in the discussion of the impacts of
NOy and SOx, and the impacts of PM2 5.
In addition to nitrogen (N) and sulfur (S) and their transformation products, other PM
components such as trace metals and organics are deposited to ecosystems and may
subsequently impact biota. Evidence for effects of PM on ecological receptors include
direct effects of airborne PM on radiative flux and both direct and indirect effects of
deposited particles. Direct effects include alteration of leaf processes from deposition of
PM ("dust") to vegetative surfaces (U.S. EPA. 2009a). Indirect effects encompass
physiological responses associated with uptake of PM components and alterations to
ecosystem structure and function (see Appendix 15).
Much of the discussion in sections dealing with chemistry, measurement, and deposition
(both wet and dry) focus on sulfuric acid (H2SO4) and HNO3, which have been long
established as the major species contributing to acid rain. Other N and S species that
either hydrolyze to form acids are also included, along with organic acids, to the extent
that they contribute substantively to acidification of terrestrial (see Appendix 5) and
aquatic environments (see Appendix S) and/or N enrichment (see Appendix 6.
Appendix 9. Appendix 10. and Appendix 11).
Major sources of the precursors (NOx, SO2) to the formation of HNO3 and H2SO4 include
on- and off-highway vehicles and electricity-generating units (EGUs). SO2 is oxidized to
H2SO4 either in the gas phase or cloud water by several well-known mechanisms. NO2 is
oxidized to HNO3, which can either deposit as HNO3 or interact with NH3 to form
particulate NH4NO3. NH4NO3 can exhibit semivolatile behavior that can substantially
alter the distance over which NH3 and HNO3 can travel. Reduced organic nitrogen
species, which could have large agricultural sources, can constitute a substantial fraction
2-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
of N delivered to ecosystems by precipitation. In areas where the rainwater pH is greater
than about 4.5, pH is not exclusively controlled by N and S species, as organic acids such
as formic, acetic, and oxalic acids can be major contributors to the acidity of rainwater.
A wide variety of N containing compounds, consisting of oxidized and reduced organic
and inorganic species contribute to wet and dry deposition. In general, deposition of
reduced (organic + inorganic) N exceeds that of oxidized N across the continental U.S.
(CONUS). Nationwide, deposition of N species occurs mainly by dry deposition of
HNO3 and NH3. The pattern is more complex for S in that for large areas, mainly in the
central U.S., wet deposition tends to dominate over dry deposition. Dry deposition of
particulate SO42 is a minor source of S to the surface, largely due to the low deposition
velocities of fine-mode particles.
Precipitation chemistry has been monitored at a large number of sites across the U.S. for
several decades as part of the National Acid Deposition Program (NADP).
Concentrations of inorganic gas and particulate phase N and S species have been
monitored across the U.S. since 1990 by the Clean Air Status and Trends Network
(CASTNET). These concentrations are then used to infer dry deposition (i.e., the transfer
of gaseous and particulate pollutants from the atmosphere to the surface by impaction
through turbulent motions and gravitational settling). Estimates of dry deposition over the
CONUS are inferred by atmospheric models. Starting in 2007, monitoring of NH3 was
initiated at a subset of CASTNET sites. Cloud deposition, which can account for the bulk
of deposition at high elevations in mountainous areas, is monitored at two locations on a
regular basis, but has been the subject of shorter term field studies in various locations in
the U.S.
Although the pH of rainwater has increased noticeably across the U.S., coincident with
notable decreases in the wet deposition of nitrate and sulfate since 1990, there are still
widespread areas, mainly in the eastern U.S., affected by acid precipitation. Deposition of
total nitrogen has not reflected the continuing decrease in NOx emissions, largely because
emissions and concentrations of NH3 have increased in many areas. Large areas, at least
one-third of the continental U.S. (CONUS), are estimated to receive at least 10 kg/ha/yr
wet + dry deposition of reactive nitrogen species. This estimate is likely too low because
reduced organic nitrogen species are not measured by the routine monitoring networks or
considered in air quality models such as U.S. EPA's Community Multiscale Air Quality
(CMAQ) modeling system.
Three of the four major contributors to inorganic N deposition are included in the
definitions of either oxides of nitrogen or particulate matter: HNO3 is an oxide of
nitrogen, NH4+ is a PM component, particulate NO3 is both a PM component and an
oxide of nitrogen. The fourth major contributor, NH3, is neither an oxide of nitrogen nor a
2-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
component of particulate matter. In a recent comparison, the contribution of NH3 to total
inorganic N deposition ranged from 19% in locations in the Northwest U.S. to 63% in
locations in the Southwest U.S., and was generally higher in summer than in winter (Li et
al.. 2016d). In general, the majority of inorganic N deposition was accounted for by
oxides of nitrogen and particulate matter in the Northwest, Northeast, Southeast, and
Rocky Mountains, but the contribution of oxides of nitrogen and particulate matter was
roughly equivalent to contributions from NH3 in the Upper Midwest, Florida, and smaller
in the Southwest (Li et al.. 2016d). However, since NH3 is a PM precursor, it can also be
definitively stated that inorganic N deposition is entirely accounted for by oxides of
nitrogen, PM components, and PM precursors.
Since the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur—Ecological Criteria (hereafter referred to as the 2008 ISA) (U.S. EPA. 2008a).
there have been a number of new developments. These apply to methods, such as
data-model fusion to integrate information for deposition across the U.S.; the use of
chemistry-transport models linking deposition to ambient air quality; the expansion of
CASTNET monitoring to include NH3 and NOy at selected sites and intercomparisons of
monitoring methods with research grade instruments; and advances being made in
satellite-based measurements in conjunction with chemistry-transport model simulations
of tropospheric NO2, SO2, and NH3 that will allow mapping of dry deposition over
remote areas with spatial resolution of -10 km / 10 km. These new developments are
described in the following sections of this appendix. Appendix 22. considers sources and
emissions of N, S, and PM to the atmosphere. Appendix 23. summarizes the atmospheric
chemical transformations of N and S compounds and formation of PM that occur during
transport from sources to deposition to the surface. Appendix 2 A describes measurement
of relevant atmospheric species, including national monitoring networks and methods.
Appendix 25. discusses the use of chemical transport models to estimate deposition.
Appendix 2J) shows the geographic distributions of atmospheric concentrations and
deposition of N, S, and PM.
2.2. Sources of Nitrogen and Sulfur Compounds and Particulate
Matter to the Atmosphere
This section describers advances in our understanding of NOx, NH3, and SO2 emissions.
Appendix 2.2.1 describes annual national emissions of each species based mainly on the
2014 National Emissions Inventory (NEI). Appendix 2.2.2 describes methods of
estimating emissions. Methods for major sources of SO2 and NOx are reliable and have
remained largely the same since the last review. The same is true to an extent for direct
PM emissions. However, there have been fundamental changes in methods for estimating
June 2018
2-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
NH3 emissions, and these are described in some detail. Appendix 2.2.3 describes
emission uncertainties, including recent comparisons between NEI data and alternative
methods of estimating emissions. Geographic distributions of emissions are presented in
Appendix 2.5. where they can be more directly compared with concentration and
deposition data.
2.2.1. National Emissions by Source
N and S containing compounds contributing to deposition can be either primary
(i.e., directly emitted from sources) or secondary (i.e., produced from atmospheric
reactions involving precursor species directly emitted from sources). Major primary
species include NOx, SO2, NH3, and reduced organic nitrogen (RON).
Table 2-1 shows nationwide emissions of NOx, SO2, and NH3 by source category
compiled from the 2014 National Emissions Inventory (NEI) and other sources.
Emissions estimates are not available for RON. For the most part, NOx, SO2, and NH3
are each emitted by different sources. NOx emissions come from several important
sources. Highway vehicles are the largest source category of NOx emissions nationwide,
but off-highway vehicles, EGU's, other stationary fuel combustion, industrial processes,
fires, and biogenic emissions from soil are all substantial contributors to total NOx
emissions. Lightning is not included in the NEI, but can also contribute substantially to
total NOx emissions. Although lightning is shown as a relatively modest source of NOx,
most production by lightning occurs during the summer and is highest in the
south-central and southwestern U.S. (Zhang et al.. 2012a).
NH3 originates mainly from agriculture, which accounts for -80% of its emissions
nationwide. Agriculturally related sources consist of livestock, including confined animal
feeding operations, and soils after addition of N containing fertilizers. Fertilizer
application occurs mainly during spring and summer. In addition to NH3, reduced organic
compounds such as urea and a wide range of proteins and other biological components
are also emitted as the result of agricultural activity. Xing et al. (2013) observed that in
contrast to SO2, NOx, and other pollutants, total national emissions of NH3 increased
from 1990 to 2010. The authors attributed this to agricultural emissions, including
livestock, which they also identified as the dominant source of NH3 emissions in the
continental U.S. However, regionally the relative importance of agricultural and vehicle
emissions is likely to be variable. The deposition of reduced nitrogen can be three times
higher near roads (Bettez et al.. 2013). and motor vehicles can be a substantial contributor
to total NH3 emissions in urban areas (Baum et al.. 2001).
June 2018
2-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Table 2-1 Emissions of NOx (nitric oxide + nitrogen dioxide), sulfur dioxide,
and ammonia by source category for 2014 (Teragrams3 N, S/yr).

NOx
so2
NHs
Highway vehicles
1
0
0.08
Off-highway
0.7
0.05 <0.01
Utilities (fuel combustion)
0.5
2
0.02
Other stationary (fuel combustion)
0.5
0.4
0.06
Industrial and other processes
0.4
0.3
0.08
Agriculture
0
0
2
Fires: wild, prescribed and agricultural
0.5
0.09
0.2
Biogenic
0.3
	b
0.08c
Total
4.0
2.3
2.3
N = nitrogen; NH3 = ammonia; NOx = the sum of nitric oxide and nitrogen dioxide; S = sulfur; S02 = sulfur dioxide; yr
a1 Teragram = 1 x 109 kg.
BNot applicable.
°Bouwman et al. (1997).
Source: httDs://www.eDa.aov/air-emissions-inventories/air-Dollutant-emissions-trends-data update except as noted.
= year.
SO2 emissions are dominated by stationary sources burning fossil fuels, particularly
EGUs, which contribute about 70% of total nationwide SO2 emissions. SO2 emissions
densities in most counties east of the Mississippi River are larger than in most counties in
the West (see Appendix 2.6.5).
Emissions of NOx and SO2 have decreased appreciably in recent years. National
emissions of NOx have decreased by 48%, and national SO2 emissions between 2002 and
2014 decreased by 68% (OAOPS-Emissions Inventory and Analysis Group. 2016).
Further details of declining emissions for these species can be found in the ISA's for
health effects for NOx and SO2 (U.S. EPA. 2008b. c). In contrast, nationwide primary
PM2 5 and PM10 emissions estimates have changed little between the 2002 NEI (U.S.
EPA. 2009a') and the 2014 NEI, with national PM2 5 emissions estimates decreasing from
5.4 to 5.3 MMT, and PM10 emissions estimates decreasing from 19.9 to 17.0 MMT.
Not included in Table 2-1 are primary emissions of PM2 5 and PM10. Nationwide PM10
emissions according to the 2014 NEI totaled 17.0 MMT. This is somewhat higher than
the SO2, NOx and NH3 emissions in Table 2-1. but 85% of these were from dust and
June 2018
2-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
fires, dominated by soil and organic matter. By comparison, total nationwide PM2 5
primary emissions totaled 5.3 MMT based on the 2014 NEI. This is comparable to the
totals for NOx, SO2 in Table 2-1 when it is considered that emissions in Table 2-1 are
given in terms of mass of N and S and does not include oxygen, while PM2 5 emissions
are based on total mass. However, primary PM2 5 emissions are also dominated by dust
(i.e., agricultural dust and road dust) and fires (i.e., wildfires, prescribed fires, and
agricultural fires), which together account for two-thirds of total nationwide PM2 5
primary emissions. PM2 5 from these source categories are mainly crustal material (dust)
and organic matter (fires).
As described in Appendix 2J_, in rural and remote areas secondary PM2 5 formed from
NOx and SO2 account for a greater fraction of PM2 5 than primary PM2 5. The fraction of
PM2 5 accounted for by NO3 and SO42 formed from SO2 and NOx in various U.S.
locations is discussed in Appendix 23. The NOx, SO2, and NHx emissions listed in
Table 2-1 cannot be used to quantitatively estimate the amount of secondary PM formed
because the precursors are not completely converted to PM. However, they provide not
only an estimate of emissions that lead to total NOy and SOx, but also provide an
estimate of the emissions that can be used in conjunction with atmospheric models to
estimate PM2 5 concentrations (See Appendix 2.5). The emissions estimated in Table 2-1
are ultimately responsible for the majority of PM2 5 in many areas, and of the fraction of
PM mass (i.e., SO42 , NO3 . NH4 ) for which ecosystem impacts are best understood.
2.2.2. Methods of Estimating Emissions
The source categories used in Table 2-1 represent groups of similar NEI source sectors.
Highway Vehicles comprise all on-road vehicles, including light-duty as well as
heavy-duty vehicles, both gasoline and diesel powered. Off-Highway vehicles and
engines include aircraft, commercial marine vessels, locomotives, and nonroad
equipment. Utilities (Fuel Combustion), also identified as electric power generating units
(EGUs), are mostly coal burning, but some facilities burn natural gas and other fuels.
Other Stationary (Fuel Combustion) includes commercial/institutional, industrial, and
residential combustion of biomass, coal, natural gas, oil, and other fuels. Industrial and
Other Processes include a variety of different industries, including chemical
manufacturing, cement manufacturing, and oil and gas production. The other processes
included in this category include gasoline stations and terminals, commercial cooking,
road and construction dust, solvent use, and waste disposal. Agriculture includes both
fertilizer application and livestock waste emissions. Fires include wildfires, prescribed
fires, and agricultural field burning. The biogenic category includes emissions from
June 2018
2-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
vegetation and soil. Both nitrifying and denitrifying organisms in the soil can produce
NOx, mainly in the form of NO.
Emissions data for each source category listed in Table 2-1 are from the 2014 National
Emissions Inventory (NEI), Version 1. The NEI is a national compilation of criteria air
pollutant and hazardous air pollutant emissions. The process of estimating emissions is
explained for each source in a detailed technical support document for the inventory
(U.S. EPA. 2016b). The NEI is maintained to support the NAAQS, and the Clean Air Act
requires states to submit emissions to the U.S. EPA as part of their State Implementation
Plans (SIPs). The Air Emissions Reporting Rule (AERR) requires agencies to report all
sources of emissions, except fires and biogenic sources. Reporting of open fire sources,
such as wildfires, is encouraged, but not required. Data in the NEI come from a variety of
sources. The emission values are predominantly from state, local, and tribal agencies and
are used wherever they are available, unless there are gaps or problems with submitted
data. U.S. EPA quality assures and augments the data provided by states to assist with
data completeness using separate augmentation procedures for each source as described
in detail in a technical support document to fill in gaps for sources and/or pollutants that
are often not reported by state, local, and tribal agencies. The intent is to create the most
complete inventory for use in air quality modeling, national rule assessments,
international reporting, and other reports. QA procedures and acceptance criteria are
detailed in the NEI technical support document (U.S. EPA. 2016b).
For nonpoint sources, U.S. EPA provides tools that state, local, and tribal agency staff
can use to generate emission estimates. For the 2014 NEI, the U.S. EPA developed
emission estimates for many nonpoint sectors in collaboration with a consortium of
inventory developers from various state agencies regional planning organizations called
the Nonpoint Method Advisory (NOMAD) Committee. More detailed NOMAD
subcommittees were established to collaborate on methods and emission factors for key
nonpoint source categories/sectors, including oil and gas exploration and production,
residential wood combustion, agricultural NH3 sources (including fertilizer and
livestock), and industrial and commercial/institutional fuel combustion, among other
sources. The U.S. EPA also generates emission estimates as stand-alone data sets
covering biogenics, agricultural livestock, fertilizer application, nonroad mobile sources,
rail emissions, and commercial marine vessel ports and in-transit (underway) sources.
U.S. EPA data sets for sources and pollutants are used only for sources not provided by
state, local, and tribal data. Tools and methods for estimating emissions from a given
source, including EGU's, agricultural livestock, fertilizer application, mobile sources,
agricultural and wildland fires (wildfires + prescribed fires), and wood combustion, were
described in the 2014 NEI technical support document (U.S. EPA. 2016b).
2-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
The 2014 NEI technical support document provides considerable detail on emission
factors and emission estimation methods by source used to generate the data in Table 2-1
(U.S. EPA. 2016b). Methods for estimating PM2 5 and PM10 emissions from dust and fire,
the two largest national sources, are derived from experimental emission factors along
with source specific information (e.g., crop type and tilling frequency for agricultural
dust, vehicle weight and miles traveled for unpaved road dust) using source specific
equations available in the NEI technical support document. These methods are largely
well-established, although they have been updated to accommodate satellite data and
emissions modeling improvements, particularly in the case of fire emissions.
Methods for estimating emissions from electric power generating units and mobile
sources, the largest sources of SO2 and NOx, are also well established, and emissions
from these sources are decreasing. In contrast, ammonia emissions in the U.S. are
increasing (Butler et al.. 2016). and significant uncertainties in the magnitude as well as
spatial and temporal variability of NH3 emissions estimates were reported in the 2008
ISA (U.S. EPA. 2008a). Two new methods for estimating ammonia emissions from
fertilizer applications and livestock waste are highly relevant to understanding NHx
sources and deposition are provided here as examples, but a similar level of detail is
given in the NEI technical support document for other sources of NH3, SO2, and NOx,
and a thorough reading of that document is necessary for a full description of emissions
estimation methods used to construct Table 2-1 (U.S. EPA. 2016b).
Soil and fertilizer emissions are treated differently in the NEI for NH3 and NOx. For NH3,
fertilizer application is recognized as a major source for which emissions are specifically
estimated, and emissions from fertilizer application are estimated only for NH3. The
approach to calculating emissions from fertilizer application in the 2014 NEI is a
completely new methodology to estimate ammonia (NH3) emissions from agricultural
soils. The approach to estimate 2014 fertilizer emissions consists of these steps: (1) run
the Fertilizer Emissions Scenario Tool for CMAQ FEST-C (vl.2)
(https://www.cmascenter.org/fest-c/) and the bidirectional version of CMAQ (v5.0.2)
(https://www.cmascenter.org/) to produce Year 2011 nitrate (NO3), ammonium (NH4 .
including urea), and organic (manure) nitrogen (N) fertilizer usage estimates, and gaseous
ammonia NH3 emission estimates respectively; (2) calculate county-level emission
factors for 2011 as the ratio of bidirectional CMAQ NH3 fertilizer emissions to FEST-C
total N fertilizer application; (3) run FEST-C to produce Year 2014 NO3, NH4 (including
Urea), and organic (manure) nitrogen fertilizer usage estimates; and (4) multiply
county-level 2014 FEST-C total fertilizer estimates by the 2011 emission factors to
estimate 2014 NH3 emissions. FEST-C reads land use data from the Biogenic Emissions
Land Use Dataset (BELD) Version 4, meteorological variables from the Weather
Research and Forecasting (WRF v3.7.1) model (http://www.wrf-model.org/index.php)
2-10
DRAFT: Do Not Cite or Quote

-------
1	and nitrogen deposition data from a previous or historical average CMAQ simulation.
2	The Environmental Policy Integrated Climate (EPIC) modeling system
3	(http: //epicapex.tamu .edu/) provides information regarding fertilizer timing, composition,
4	application method, and amount. Figure 2-1 provides a comprehensive flowchart of the
5	complete EPIC/FEST-C/WRF modeling system.
areNEI emis
The Fertilizer Emission Scenario Tool for CMAQ
(FEST-C)
Source: U.S. EPA (2016b).
Figure 2-1 Modeling system used to compute 2014 Fertilizer Application
Emissions.
6	Fertilizer application can also lead to NOx emissions, but it does not dominate soil
7	emissions of NO. Biogenic emissions of NO are computed based on 2014 meteorology
8	data from the Weather Research and Forecasting (WRF) model version 3.8 (WRFv3.8)
9	and using the Biogenic Emission Inventory System, Version 3.61 (BEIS3.61) model,
10	based on land use and vegetation data (U.S. EPA. 2016b). The contribution of fertilizers
June 2018	2-11	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
to soil NOx emissions is not estimated in the NEI, but it has been estimated as 10%
globally (Hudman et al.. 2012). Biogenic emissions of NH3 are not estimated in the NEI,
but aside from fertilizer application it is a minor contributor, as shown in Table 2-1.
Further details on estimating biogenic NOx emissions are given in the NEI Technical
Support Document (U.S. EPA. 2016b).
Livestock waste is another important source of ammonia in the U.S. In the 2014 NEI, the
U.S. EPA has updated the methodology for ammonia emissions from the
housing/grazing, storage and application of manure from beef cattle, dairy cattle, swine,
broiler chicken, and layer chicken production. Cows, swine, and chickens account for
95% of national NH3 emissions from livestock waste in 2014. The approach to estimate
2014 livestock NH3 emissions from these animals consists of these general steps:
(1) estimate 2014 county-level animal populations using 2012 and 2014 USDA
agricultural census data; (2) use a model developed by Carnegie-Mellon University
(Mcquilling and Adams. 2015; Pinder et al.. 2004a; Pinder et al.. 2004b) to produce daily
resolved, climate-level emission factors for a particular distribution of management
practices for each county and animal type, as expressed as emissions/animal; and
(3) multiply the county animal populations by the daily emission factor for each county
and animal type to estimate emissions per day and sum daily emissions across the entire
year for each county and SCC to produce annual emissions for use in the NEI. The model
inputs and outputs are shown in Figure 2-2.
Source: U.S. EPA (2016b).
Figure 2-2 Process to produce specific location and practice specific daily
emission factors for livestock waste.
June 2018
2-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Similar details are given for estimating emissions for major sources of NOx and SO2 in
the NEI technical support document (U.S. EPA. 2016b). including EGU's, on-road
mobile sources, marine vessels, locomotives, other nonroad sources, airports, rail yards,
landfills, agricultural and wildland fires, wood combustion, other fuel residential and
industrial fuel combustion, charcoal grilling, waste disposal, vegetation and soil, and
other sources. Air emissions data from the 2014 Toxic Release Inventory (TRI) were also
used in the 2014 NEI to supplement point source NH3 emissions provided to the U.S.
EPA by state, local, and tribal agencies. The TRI is a U.S. EPA database containing data
on disposal or other releases including air emissions of over 650 toxic chemicals from
approximately 21,000 facilities. Data are submitted annually by U.S. facilities that meet
TRI reporting criteria.
2.2.3. Evaluation and Uncertainty
As described in the 2008 ISA (U.S. EPA. 2008a'). emissions from different sources in the
NEI are estimated with a wide range of methods that include direct measurements,
indirect measurements, model predictions, and assumptions. Because there are unknown,
incomplete, and variable emission rates, as well as unknown sources that are not
represented, the NEI reflects an on-going process of updating increasing or declining
emissions, improving estimation methods, and filling data gaps as measurements become
available or understanding of emissions changes. Often, steps are taken to reduce errors
in estimation as they are discovered, resulting in improved estimates as uncertainties are
found. For example, the estimate of Gilliland et al. (2003) that annual NEI NH3 was 37%
higher than estimates based on modeled NH4 + deposition led to the development of lower
emission factors for nondairy cows and swine (U.S. EPA. 2008a'). Because of these
unknowns and limitations, quantitative uncertainty estimates based on probability density
functions or other statistical methods are not provided with NEI data. Instead, emission
factors receive a rating based on the reliability of methods used for determining the
emission factor. The rating is based on both the number of representative sources and the
characteristics of the data used to determine the emission factor, but it does not imply
statistical error bounds or confidence intervals for the emission factor (U.S. EPA. 1996).
As an alternative, uncertainties are often evaluated through separate efforts using a
variety of techniques. These techniques include comparing inventory predictions with
measured long-term trends, comparing emission estimates derived from principle
component analysis or other statistical methods, comparing emissions estimated by
inverse modeling of chemical transport models, and comparison with satellite data (U.S.
EPA. 2008a). The distinction between the inventory compilations like the NEI and
alternative satellite- and model-based methods of estimating emissions is generally
June 2018
2-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
described in terms of bottom-up and top-down estimates. The entries in emissions
inventories are obtained using a bottom-up approach, in which entries are based on
emissions factors, activity rates, and control device efficiency for various source types.
This contrasts to a top-down approach in which measurements of pollutant concentrations
from satellites, aircraft, or surface monitors are used to constrain a priori estimates of
emissions using a chemistry-transport model (CTM).
Because of this variety of top-down approaches and the number of separate studies
resulting in a wide range of estimated uncertainties, there is no single estimate of
uncertainty that applies to either total emissions or emissions from individual sources in
Table 2-1. However, reports from numerous publications on emissions inventory
evaluation have resulted in a wide range of uncertainty estimates for application to NEI
data, and these were summarized in the 2008 ISA (U.S. EPA. 2008a). Across all sources,
total NOx emissions estimates from satellite data ranged from "highly consistent" with
NEI estimates (Martin et al.. 2006) to 68% higher than NEI estimates (Jacglc et al..
2005). Fewer estimates of individual source emissions were evaluated, with NOx
emissions both higher and lower than estimates using other methods.
Some recent work has shown summertime over-prediction of model NOx estimates using
recent U.S. EPA inventories (e.g., 2008 and 2011) when compared against monitored
ambient concentrations (Cantv et al.. 2015; Anderson et al.. 2014a). Appel et al. (2017)
found in their simulations that NOx model over-estimates in summer were greatly
diminished or reversed in other seasons. Anderson et al. (2014a) and Travis et al. (2016)
concluded that emissions of NOx from mobile sources are being overestimated and are
the source of this bias. Studies over Texas (Souri et al.. 2016; Tang et al.. 2015)
suggested a smaller or no bias in on road NOx emissions compared to Anderson et al.
(2014a) and Travis et al. (2016). and suggested that both high and low biases in other
NOx source categories (e.g., area sources, point sources, soil NO) impact
model/measurement discrepancies. Marr et al. (2013) used near-road measurements to
also conclude that mobile source NOx emissions in U.S. EPA's 2008 NEI agreed well
with measurements (i.e., within 3%). The cause of discrepancies between measured and
modeled concentrations are difficult to diagnose because the emission modeling process
and associated photochemical modeling is complex. Researchers are continuing to
investigate this question.
There can be higher uncertainties for specific sources. For example, about 60% of the
total NOx emitted by soils nationwide is estimated to occur in the central Corn Belt of the
U.S. Spatial and temporal variability in soil NOx emissions can lead to uncertainty in
emissions estimates. Soil emissions occur mainly during summer and across the entire
country, including areas where anthropogenic emissions are low. Emission rates depend
2-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
primarily on fertilization amount, soil temperature, and moisture. Models of NOx
emissions from soils (e.g., (Hudman et al.. 2012) include these dependencies, but most
measurements on which they are based are made at temperatures <30°C. However, in
agricultural areas subjected to very high temperatures (>40°C) like the Imperial Valley,
CA, emissions factors for NO following fertilizer application ranged from 1.8 to 6.6%, as
compared to estimates of typically ~1 to 2% (Oikawa et al.. 2015). Oikawa et al. (2015)
also suggest that in many areas of the Southwest, the NEI overestimates anthropogenic
emissions at the expense of soil emissions and that these soil emissions have a noticeable
effect on ozone formation. Travis et al. (2016) estimate that combustion accounts for
68% of NOx emissions in the Southeast in summer, with the remainder from soils. These
results indicate that soil emissions need to be better understood. Estimating emissions
from highway vehicles can also be challenging because there is a wide variation in
emissions between different vehicles.
Activity rates and uncertainties for NH3 are difficult to quantify, and estimates have yet
to be made for reduced organic nitrogen. NH3 emission estimates are generally more
uncertain than NOx and SO2 emission estimates because of the variety of agricultural
practices used, re-emission after deposition, and the dispersed nature of agricultural
processes, as well as the complex influences of meteorology on processes controlling
transformation and removal of nitrogen species on spatial and temporal emission patterns.
As described in Appendix 2.5.2. NO2 and NH3 can be both emitted from and deposited to
soils, water, or vegetation depending on their atmospheric concentrations and
characteristics of the underlying surface.
Mixed results were reported in evaluating the total NH3 emissions estimates in the 2008
ISA (U.S. EPA. 2008a). However, as described above, there have been major changes in
estimation methods for the most important ammonia sources since that report, and more
recently Paulot et al. (2014) showed that a range of published bottom-up and top-down
estimates of annual total U.S. ammonia emissions agreed within 10%. The NEI national
estimate of 2.3 Tg NH3 from Table 2-1 was within 20% of the average across all
estimates of 2.8 ± 0.2 Tg NH3 reported by Paulot etal. (2014). However, in the same
comparison there was a divergence in the timing of the seasonal maximum, and
agreement varied considerably temporally and spatially (Paulot et al.. 2014).
Most SO2 emissions originate from point sources having well-known locations and
identifiable fuel streams. Uncertainties in annual emissions were estimated to range from
4 to 9% for SO2 and slightly larger for NOx from the same point sources identified in the
1985 NAPAP inventories for the U.S. (Placet et al.. 1990).
2-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
2.3.
Atmospheric Chemistry of Nitrogen and Sulfur Species and
Particulate Matter (PM)
The atmospheric chemistry of N and S species relevant for the production of ecosystem
nutrients and acidic species was extensively reviewed in the 2008 ISA (U.S. EPA.
2008a). The main findings from that review and key findings from more recent studies
and reviews are included here. Appendix 2.3.1 describes atmospheric NOx chemistry and
the formation of HNO3. Appendix 2.3.2 describes atmospheric sulfur oxide chemistry and
the formation of H2SO4. Appendix 2.3.3 reviews the role of ammonia as the most
important atmospheric base for neutralizing atmospheric nitric and sulfuric acids and
forming PM. The chemistry of all of these species largely controls the extent of acid
deposition as well as the fraction of nitrogen in particulate matter, which in turn
determines deposition rate and transport distance. The remaining sections review
atmospheric organic sulfur and nitrogen compounds (Appendix 2.3.4). atmospheric
organic acids (Appendix 2.3.5). and formation of PM25 (Appendix 2.3.6).
2.3.1. Nitrogen Oxides
NOx (NO + NO2) is the precursor for oxidized nitrogen species that contribute to acidic
deposition. More specific details on the chemistry and transformation of NOx can be
found in the 2016 ISA for Oxides of Nitrogen—Health Criteria (U.S. EPA. 2016f) and
for SOx in the 2008 ISA for Sulfur Oxides (U.S. EPA. 2008c). Hence, those topics are
only briefly recounted here with special reference to the secondary NOx and SOx
NAAQS. Oxidized nitrogen species (NOy) are introduced into the atmosphere as NOx,
mainly from fossil fuel combustion as described in Appendix 2^2. Figure 2-3 summarizes
the atmospheric reactions of NOx, showing rapid inter-conversion of NO and NO2 in
sunlight, with slower formation of more oxidized organic and inorganic products (NOz).
A large number of oxidized nitrogen species in the atmosphere are formed from the
oxidation of NO and NO2 (shown in the inner box). These include nitrate radicals (NO3),
nitrous acid (HONO), nitric acid (HNO3), dinitrogen pentoxide (N2O5), nitryl chloride
(CINO2), peroxynitric acid (HNO4), peroxyacetyl nitrate (PAN) and its homologues
(PANs), other organic nitrates, such as isoprene- and monoterpene-derived nitrates, and
particulate nitrate (pNOs ). These species (and NH3) are characterized by large
differences in their solubilities (Table 2-2). which determine their ability to be taken up
by cloud droplets, airborne particles, and moist surfaces.
June 2018
2-16
DRAFT: Do Not Cite or Quote

-------
emissions
Ca2+ = calcium ion; HN03 = nitric acid; H02 = hydroperoxy radicals; hv = solar photon; Mg2+ = magnesium; NH3 = ammonia;
NH4+ = ammonium; NO = nitric oxide; N02 = nitrogen dioxide; N03" = nitrate ion; NOx = the sum of NO and N02; NOz = oxidation
products of NOx; NOY = NOx + NOz; 03 = ozone; PAN = peroxyacetyl nitrate; R02 = organic peroxy radicals.
Note: The inner shaded box contains NOx (NO + N02). The outer box contains other species (NOz) formed from reactions of NOx.
All species shown in the outer and inner boxes are collectively referred to as NOY by the atmospheric sciences community.
Source: NCEA.
Figure 2-3 Schematic diagram showing pathways for reactive nitrogen
species in ambient air.
June 2018
2-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
Table 2-2
Henry's law coefficients for selected reactive nitrogen species at
25°C in water.

Compound
Coefficient (mol/kg/bar)
HNOs

2.6 x 10®
HONO

49
NO

0.0019
NO2

0.012
PAN

4.1
NHs

61
1 bar = 105 Pa; C = Celsius; HN03 = nitric acid; HONO = nitrous acid; NH3 = ammonia; NO = nitric oxide; N02 = nitrogen dioxide;
PAN = peroxyacetyl nitrate.
Source: adapted from Sutton et al. (2011).
Reactions producing more oxidized forms of nitrogen (NOz) involve mainly O3, OH, and
organic radicals with NO and NO2. The reaction of NO2 with OH leads directly to HNO3:
JVO2 + OH + M HNO3
Equation 2-1
The reaction ofN02 with O3 produces nitrate radical (NO3), which reacts further to form
dinitrogen pentoxide (N2O5), and ultimately also produces HNO3:
1VO2 + O3 NO3 + O2
Equation 2-2
NO2 + NO3 N2O5 (equilibrium)
Equation 2-3
N2O5 + H2O -> 2HN03
Equation 2-4
The relative importance of these two paths for producing HNO3 is strongly location and
seasonally dependent, with the first path dominating when OH radicals are abundant
(during the day) and the second during the night and under cold conditions. Warneck
(1999) estimated that most HNO3 is formed in the sunlit portions of clouds by the
reaction of NO2 with OH, with much smaller amounts from the pathway involving N2O5
hydrolysis. Because it is highly soluble, HNO3 is taken up by particles or cloud droplets
to form N03 and is also deposited onto moist surfaces, such as on vegetation. HNO3 also
June 2018
2-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
recycles back to NO2 in the gas phase by photolysis and reaction with OH radicals, but on
timescales longer than that for uptake by cloud droplets, particles, and the surface.
Whereas photolysis of HNO3 is slow (x ~ 106 s) in the gas phase, it can be two orders of
magnitude faster on moist surfaces (Ye et al.. 2016). releasing NO2 and/or HONO back to
the atmosphere.
NO2 reacts with organic peroxy radicals to form organic nitrates such as peroxyacetyl
nitrate (PAN) and its homologues as shown on the right side of Figure 2-3; other RO2NO2
compounds are much less stable than PANs. NO and NO3 radicals also react with organic
radicals produced by the oxidation of isoprene and monoterpenes to form a wide range of
organic nitrates. Considering the troposphere as a whole, most of the mass of NOz shown
in Figure 2-3 is in the form of PAN and HNO3. However, organic nitrates such as
isoprene- and monoterpene-derived nitrates increase in importance in the planetary
boundary layer (PBL), and are likely to be dominant in vegetated areas (Kim et al..
2015a; Min et al.. 2014).
In forested areas, the initial step in the production of isoprene nitrates (INs) is most often
the reaction of isoprene with OH radicals to produce isoprene peroxy radicals. These can
react with HO2 radicals, other RO2 radicals, or isomerize to produce a variety of organic
compounds. They can also react with NO to produce multifunctional organic nitrates.
Lifetimes on the order of one to a few hours can be estimated for these first generation
INs based on their reactions with OH radicals and O3 (Lockwood et al.. 2010; Paulot et
al.. 2009). The reaction products can further react with NO (after internal rearrangement)
to form secondary organic nitrates such as ethanal nitrate, methacrolein nitrate,
propanone nitrate, and methyl vinyl ketone nitrate. The second-generation organic
nitrates are more stable than the first-generation INs because they lack a double carbon
(C = C) bond. Obviously, the relative importance of pathways forming nitrates or other
products depends on the ambient concentrations of NO and other oxides of nitrogen for
which many key experimental details are still lacking. During the SEAC4RS
measurement campaign, which took place in the summer of 2013 in the southeastern
U.S., Travis et al. (2016) found that these two pathways were of comparable importance.
In addition to oxidation initiated by OH radicals, isoprene is also oxidized by NO3
radicals. Rollins et al. (2009) determined a yield of first-generation carbonyl nitrates of
70% based on experiments in large reaction chambers. These first-generation nitrates can
further react with NO, leading to the production of second-generation organic (alkyl)
nitrates. Mao et al. (2013) estimated that the global mean lifetime is ~5 days for these
second-generation organic nitrates. Mao et al. (2013) also suggested that the export of
INs and other organic nitrates followed by their decomposition is potentially a larger
source of NOx to the boundary layer of the western North Atlantic Ocean than the export
2-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
of PANs. Some INs are low enough in volatility that they can partition to the particle
phase [e.g., (Rollins et al.. 2009)1. Once in the particle phase, the INs hydrolyze to form
HNO3 and an alcohol, with a rate constant that correlates strongly with the acidity of the
particles (Rindclaub et al.. 2015; Jacobs et al.. 2014).
In addition to considering the chemistry of isoprene-derived nitrates during SEAC4RS,
Fisher et al. (2016) considered the formation of organic nitrates derived from the
oxidation of monoterpenes with either one or two double bonds. Their modeling results
suggest that isoprene- and monoterpene-derived nitrates account for 25 to 50% and -10%
of total organic nitrates and that production of isoprene- and monoterpene-derived
nitrates account for -20% of the net loss of NOx emitted in the Southeast during summer.
Fisher et al. (2016) also noted that production of organic nitrates involving biogenic
VOCs is the dominant NOx sink only in areas where elevated levels of biogenic VOCs
coincide with low NOx levels (otherwise the major sink would be formation of HNO3).
As a result, these processes will represent only a minor pathway for NOx loss. In any
event, as with isoprene-derived nitrates, monoterpene-derived nitrates are also mainly
taken up by particles with formation of HNO3. Uptake by particles was estimated by
Fisher et al. (2016) to account for -60% of the removal of gas-phase organic nitrates,
with -20% recycled back to NOx and another 15% deposited to the surface.
2.3.2. Sulfur Oxides
SO2 is the only gas phase form of SOx (SO2 + SO42 ) emitted in the tropospheric
boundary layer at concentrations of concern for environmental exposures (U.S. EPA.
2008c). It reacts in both the gas phase and in aqueous solution in clouds and particles to
form SO42 . As described in the 2008 ISA (U.S. EPA. 2008a). the steps involved in
aqueous-phase oxidation of SO2 begins with dissolution of SO2 following Jacobson
(2002):
SO2(g) S02(aq)
Equation 2-5
and is followed by formation and dissociation of H2SO3:
S02(aq) + H2O H2S03(aq) H+ + HS03~ 2H+ + S032~
Equation 2-6
Dissolved SO2 thus rapidly partitions into four forms with the same oxidation state, with
their relative concentrations dependent on pH:
June 2018
2-20
DRAFT: Do Not Cite or Quote

-------
S(IV) = S02(aq) + H2S03(aq) + HS03~(aq) + S032~(aq)
Equation 2-7
1	S(IV) is then oxidized to SO42 in cloud water primarily by either H2O2, O3, and O2 in the
2	presence of dissolved Fe(III). Reaction with H2O2 is most important at pH less than about
3	5.3, and reaction with either dissolved O3 or with O2 catalyzed by Fe(III) becomes most
4	important at pH greater than about 5.3, as shown in Figure 2-4 (Seinfeld and Pandis.
5	1998).
pH
aq = aqueous; Fe(lll) = iron (oxidation number III); H202 = hydrogen peroxide; Mn(ll) = mangnanese (oxidation number II);
N02 = nitrogen dioxide; 03 = ozone; S(IV) = sulfur (oxidation number IV); S02 = sulfur dioxide.
Concentrations assumed are: [S02(g)] = 5 ppb; [N02(g)] = 1 ppb; [H202(g)] = 1 ppb; [03(g)] = 50 ppb; [Fe(l Il)(aq)] = 0. 3 |jM;
[Mn(ll)(aq)] = 0.03 |jM.
Source:Seinfeld and Pandis (1998).
Figure 2-4 Rate of conversion of sulfur (IV) to sulfur (VI) by different
oxidation paths as a function of pH.
6	The remaining SO2 is oxidized to H2SO4 in the gas phase with a characteristic timescale
7	of -10 days [based on OH = 106/cm3 and rate coefficient = 1.3 x 10 l2/cm7molcc/s:
8	(Sander et al.. 2011)1 following a multistep process:
June 2018	2-21	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
SO2 + OH + M HSOs + M
Equation 2-8
HSOs + 02 SO 3 + HO2
Equation 2-9
and/or by
502	+ sCI -> SO3 + products
Equation 2-10
where sCI is a stabilized Criegee intermediate (Berndt et al.. 2012; Mauldin et al.. 2012;
Welz et al.. 2012) and products refer to other organic radicals. Criegee radicals are
produced by the reaction of alkenes with O3 during both night and day. The relative
importance of the OH and sCI pathways depends in large measure on the local
concentration of alkenes, in particular biogenic alkenes. Welz et al. (2012) also raised the
possibility that Criegee radicals might be important for the oxidation of NO2 to form
nitrate radicals. SO3 produced by either path further reacts to form gas-phase H2SO4 via
503	+ H2O -> H2SO4
Equation 2-11
Because H2SO4 is extremely soluble, it is removed rapidly by transfer to the aqueous
phase of particles and cloud droplets.
2.3.3. Acid Neutralization by Ammonia
As the most common soluble base in the atmosphere, NH3 plays a key role in neutralizing
the acidity in ambient particles and in cloud, fog, and rainwater resulting from dissolution
of H2SO4 and HNO3, and the weak acidity due to organic acids. The atmospheric lifetime
of NH3 with respect to oxidation by OH radicals is ~2 months [based on
OH = 106 molec/cm3 and rate coefficient =1.6/ 10 ' Vcm Vmolcc/s: (Sander et al..
2011)1. As a result, uptake by cloud droplets, particles, and the surface is favored over
reaction with OH radicals. Xu and Penner (2012) estimate a globally averaged lifetime
for NH3 of ~11 hours as a result of these processes, implying strong spatial and temporal
variability of NH3 concentrations.
Sulfuric acid can be partly or totally neutralized by NH3. Seinfeld and Pandis (1998)
define two regimes: (1) ammonia poor [TA] <2 [TS] and (2) ammonia rich [TA] >2 [TS],
where TA and TS refer to total ammonia, ammonium and sulfate concentrations in gas,
aqueous, and solid forms. In the first regime, there is partial neutralization; sulfate is in
the form of (NH4)HS04, the vapor pressure of NH3 is very low, equilibrium favors
June 2018
2-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
formation of ammonium sulfate over ammonium nitrate, any nitrate is driven to the gas
phase, and ammonium nitrate levels are low (or even zero). In the second regime, sulfate
is in the form of (NH^SO^ and any NH3 left over could react with HNO3 to form
NH4NO3. However, this conceptual model neglects interactions with organic compounds.
Analysis of data by Kim et al. (2015a) from the SEAC4RS field study and the Chemical
Speciation Network (CSN) during August-September of 2013 indicates that the extent of
neutralization of sulfuric acid and acidic sulfate by ammonium was incomplete in the
Southeast despite an excess of atmospheric NH3. Kim et al. (2015a) suggested that uptake
of NH3 is inhibited by organic compounds in particles. This suggestion is in accord with
laboratory studies of Liggio et al. (2011) who found that organic compounds, especially
terpenes and «-alkanes on particle surfaces are effective in inhibiting NH3 uptake by
particles.
NH4NO3 is in thermodynamic equilibrium with gas-phase NH3 and HNO3. The
equilibrium constant is extremely sensitive to variations in relative humidity and
temperature such that it varies over several orders of magnitude depending on
atmospheric conditions, but in general, lower temperature and higher relative humidity
(e.g., during winter) shifts the equilibrium towards condensed phase NH4NO3. The effects
on phase partitioning are pronounced because of the large variation in the equilibrium
constant, K, (—10—103 ppb2) between summer and winter conditions in many locations.
Also, as noted by Malm et al. (2016). NH4NO3 can volatilize and reform multiple times
during transport away from sources of NH3 and HNO3. Because the atmospheric lifetimes
of NH3, HNO3, and NH4NO3 differ substantially from each other, local conditions of
temperature and relative humidity, by implication, control how far these species can
travel.
Although the above considerations apply to particles in general, it should be remembered
that the mass of airborne particles is present in two distinct size fractions, each with its
own characteristic composition [see U.S. EPA (2009aYI. These differences determine the
size fraction in which pNCh will be found. Because SO42 is found mainly in the fine
particle mode these considerations tend to apply more to the atmospheric fine mode.
Displacement of HC1 (and other hydrohalic acids) from marine aerosol (found typically
in the coarse mode) by gas-phase HNO3 has long been known to occur, resulting in
particulate nitrate (pNCh ) being associated with sodium in the coarse mode in many
coastal areas. Brimblecombe and Clegg (1988) provided a detailed evaluation of the
thermodynamic data and a discussion of this process. Wolff (1984) found that
coarse-mode pNCh is formed by adsorption of HNO3 on basic soil particles (i.e., those
containing Ca2+ and Mg2+). These distinctions between the behavior of pNCh in the fine
and coarse modes are important as deposition rates for these two size modes can differ
2-23
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
appreciably and there can be large differences in the ratio of fine to coarse pNC>3 as
shown in Appendix 2.5.2.1.
The composition of rainwater and of particles is strongly affected by pH. As described
above, pH determines the distribution of S(IV) species in cloud water, rainwater and the
aqueous phase of particles. This in turn impacts their oxidation processes
(Appendix 2.3.2). the solubility of trace metals, and the partitioning of weak acids among
other factors. As the most abundant base in the atmosphere, NH3 has a strong influence
on pH of cloud water. However, the role of NH3 as a base is limited to atmospheric
processes. Once deposited in soil, oxidation of NH3 and NH4 to NO3 (during
nitrification) produces an amount of H+ equivalent to HNO3 deposition rScheffe et al.
(2014) and references therein].
2.3.4. Organic Nitrogen and Sulfur
In addition to deposition of NOy and NHx, the deposition of other nitrogen compounds,
in particular dissolved organic nitrogen (DON) also occurs. Proteins, amino acids, urea,
amines, and other DON compounds can contribute to acidification in soils and be an
important source of nutrients to terrestrial and aquatic environments (Jickells et al.. 2013;
Cape etal.. 2011; Cornell. 2011b; Sutton etal.. 2011). The content of organic nitrogen in
particles and rainwater can be characterized in two ways. First, it can be calculated as the
difference between total N as measured by total elemental analysis [see e.g., Bronk et al.
(2000)1 minus NO3 and NH4+. In the second way, the content of organic nitrogen in
particles and rainwater can be characterized by measuring the concentrations of
individual species. However, the number of species constituting DON at a particular
location can be quite large. For example, Altieri et al. (2009a) detected several thousand
organic N containing species in precipitation samples collected in New Jersey and found
the overall composition was consistent with oligomerization of amino acids; in most
compounds, N was in reduced form.
Cornell (2011a) estimated based on measurements reported in 58 published studies that
organic N constitutes 35% of total N in rainwater in North America. Jickells et al. (2013)
estimated based on data from a number of measurement sites (n = 115 globally), that
average DON in rainwater contributes -25% of the flux of total nitrogen. They also
reasoned that because it is correlated with total nitrogen in rainwater (If = 0.57), which
has a large anthropogenic component, DON might also have a large anthropogenic
component. Further description of DON measured at sites in the CONUS are deferred to
Appendix 2.6.2.
June 2018
2-24
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
A large number of organosulfates (R-O-SO3H) have been detected in rainwater samples
(Alticri et al.. 200%). However, their abundances could not be determined. Tolocka and
Turpin (2012) estimated that organic sulfates could contribute up to 5 to 10% of organic
mass on average to particle mass based on measurements taken at 12 sites across the U.S.
Liao et al. (2015) found that organic sulfates accounted a few percent of particulate
sulfate, mainly from the two most abundant forms, isoprene epoxydiol sulfate and
glycolic acid sulfate. Organic sulfates such as these have high acid dissociation constants
and are expected to act as singly charged species.
Phytoplankton emit copious amounts of dimethyl sulfide, which can be oxidized to sulfur
dioxide and to methane sulfonic acid. The SO2 that is formed can then be a source of
H2SO4 in coastal areas.
2.3.5. Organic Acids
The effects of deposition of acidic sulfur and nitrogen should be considered in the context
of a more complete description of the composition of rainwater, including organic acids.
However, organic acids are not routinely measured by monitoring networks because they
are unstable with respect to microbial degradation following collection. As a result, data
for organic acids in rainwater are sparse. Formic and acetic acids are typically the most
abundant organic acids found in rainwater in the U.S. ("Willev et al.. 2011; Avery et al..
2006; Talbot etal.. 1990). They are largely secondary in origin (i.e., produced in the
atmosphere by the photochemical oxidation of biogenic and anthropogenic
hydrocarbons). Paulotetal. (2011) suggested that isoprene oxidation is the largest global
source of formic and acetic acids. These acids are also produced by the oxidation of
ethane and propylene emitted in automobile exhaust. Their abundances in rainwater and
their effects on pH are not negligible. For example, Willev et al. (2011) found that formic
acid (pKa = 3.75) and acetic acid (pKa = 4.76) were the major organic acids present and
contributed -22 and 5%, respectively, of free acidity (mean pH = 4.65) in rainwater
samples collected at Wilmington, NC in 2008. In addition, other organic acids
(e.g., oxalic acid, lactic acid) have been found to be present at much lower levels at this
site (Avery et al.. 2006). Vet et al. (2014) noted that organic acids should be monitored in
areas where the concentration of H+ is <5 (j,eq/L (or pH > 5.3). As will be seen in
Appendix 2.6. this condition is met in areas like the Northwest where concentrations of
N03 and S042 in rainwater are low. Even in areas where the effects of organic acid
neutralization by NFU+ might be small, the vapor pressures of some organic acids
(e.g., oxalic acid) would be reduced by orders of magnitude, resulting in increased uptake
of the organic acid from the gas phase and growth of particles (Ortiz-Montalvo et al..
2014; Paciga et al.. 2014).
June 2018
2-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
2.3.6.
Particulate Matter (PM)
The chemistry of NOx and SO2 described in Appendix 2.3.1 and Appendix 2.3.2. and the
neutralization reactions of NH3 described in Appendix 2.3.3 are relevant not only to
understanding the distribution of NOy and SOx species, but also in explaining a large
fraction of PM25 in most of the U.S. Figure 2-5 shows PM25 composition in numerous
U.S. locations. In all locations SO42 and NO3 account for a substantial fraction, and in
many cases the majority, of PM2 5. In general, SO42 accounts for an increasing fraction of
PM2 5 moving east or south, and NO3 for a greater fraction moving west or north.
Figure 2-5 also shows that PM2 5 concentrations are lower and SO42 accounts for a much
greater fraction of PM2 5 mass in 2003-2005 than in 2013-2015. This reflects the steep
decline in SO2 emissions over this period (Appendix 2.2.1) and demonstrates that it has
greatly impacted PM2 5 composition and concentration in the U.S.
The decrease in SO42 contribution is so large that in many locations where SO42 was the
greatest contributor to PM2 5 mass in 2003-2005, organic carbon was more abundant in
2013-2015. However, as described in Appendix 2.1. organic matter does not contribute
as much to acidification or nutrient enrichment as SO42 and NO3 . and SO42 and NO3
still account for the majority of PM2 5 mass in many locations. The remaining mass of
PM2 5 is composed of elemental carbon, sea salt (mostly Na and CI), and crustal material
(Si and A1 are most abundant elements). Monitoring methods are described and spatial
and temporal trends for PM2 5 species are further developed in the 2009 PM ISA (U.S.
EPA. 2009a). There is much less information on PM10 or PM10-2.5 composition because of
the lack of routine monitoring on the scale that has been implemented for PM2 5 species.
June 2018
2-26
DRAFT: Do Not Cite or Quote

-------
Figure 2-5 Contributions of organic carbon (OC), elemental carbon (EC),
sulfate, nitrate, sea salt, and crustal components to PM2.5 at
selected sites (A) 2003-2005 (B) 2013-2015.
June 2018
2-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
2.4.
Concentration and Deposition Measurements
An extensive review of techniques for measuring NOx, NOz, NOy, NHx, and SOx
appeared in the 2008 ISA (U.S. EPA. 2008a'). Updates to techniques for measuring NOx,
NOy, and SOx species can be found in the latest ISA for Oxides of Nitrogen (U.S. EPA.
2016f), Sulfur Oxides (U.S. EPA. 2008c'). and PM, including PM2 5, PM10, and PM10-2.5
(U.S. EPA. 2009a') Health Effects, to which the reader is referred for details. In the
following sections, measurements of NOx, SO2, and PM in national networks are only
briefly discussed and the measurement of species most relevant to acid and nutrient
deposition and measurement methods for wet and dry deposition are the main focus.
Appendix 2.4.1 explains the roles of various national and regional monitoring networks
in place to support the NAAQS or to collect data used for estimating acid and nutrient
deposition. Appendix 2.4.2. Appendix 2.4.3. and Appendix 2.4.4 describe methods used
to measure gas-phase oxides of nitrogen, reduced nitrogen, and sulfur oxides that are not
based on filter collection. Each of these sections is divided into separate discussions of
the methods used in monitoring networks, remote sensing methods, and recent advances
in research methods and other methods that are effective for intensive field studies but
impractical for routine monitoring. Satellite-based remote sensing methods are useful
because network coverage is often sparse and satellite-based measurements are becoming
a more widely used alternative to ground-based measurements. Appendix 2.4.5 describes
filter-based methods used in CASTNET and other networks for mainly particulate
species, but also for some gases, including HNO3 and SO2. Appendix 2.4.6 describes wet
and dry deposition measurement and recent advances.
2.4.1. Monitoring Networks
Federal Reference Methods (FRMs) have been established and national monitoring
networks put in place for NO2 as the indicator of oxides of nitrogen, SO2 as the indicator
of sulfur oxides, and PM2 5 and PM10 as indicators for PM. These methods and networks
are described in detail in recent Integrated Science Assessments for Health Effects of
Oxides of Nitrogen (U.S. EPA. 2016f). Sulfur Oxides (U.S. EPA. 2008c). and Particulate
Matter (U.S. EPA. 2009a'). However, in general the large fractions of N and S deposition
accounted for by species other than NO2 and SO2 make measurements of these indicator
species alone inadequate for estimating deposition amounts of total oxides of nitrogen
and total sulfur oxides. Similarly, it has long been established that wet deposition of S is
usually dominated by SO42 rather than SO2 (Dana. 1980). In this respect, PM2 5
monitoring is potentially useful because it efficiently collects the range of PM species
involved in acidification and N deposition. However, variability of SO42 and NO3 as a
June 2018
2-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
fraction of total PM2 5 presents a challenge for relating PM2 5 mass to these effects. An
additional challenge is the estimate of deposition from PM2 5 concentrations, because of
the strong dependence of deposition flux on particle size (see Appendix 2.5.2) and
general unavailability of size distribution measurements.
In principle, a multipurpose, multipollutant monitoring network could efficiently meet the
needs of estimating N and S deposition and air monitoring for ecosystem protection,
while at the same time addressing other national air monitoring priorities. Such a network
could include measurements of other N and S species besides NO2 and SO2 as well as
other species that are otherwise not routinely monitored to better understand a variety of
air pollution processes. This is the overall concept behind the National Core Network
(NCore), a newly developed multipollutant monitoring network, and measurements of
NOy and NH3 were included as NCore monitoring in part because of their relevance to
atmospheric deposition (Scheffe et aL 2009). NCore has been operating since January 1,
2011 and has 80 monitoring sites designed for measuring multiple pollutants (Wcinstock.
2012). The network provides a core of sites that measure SO2, NO2, NOy, and PM
components including ammonium, nitrate, and sulfate, but with sparser coverage than the
FRM networks for SO2 or NOx. Because NOy is measured rather than NOx, and because
of collocated SO2 and SO42 measurements, ambient concentrations of both NOy and SOx
can be determined from NCore data, so that these data can be used to estimate total
deposition of oxides of nitrogen and sulfur. However, because of the wide range of
deposition velocities for different species, NOy measurements alone are not sufficient for
estimating deposition and species concentrations are also necessary. A further
disadvantage is that most NCore sites are located in urban areas.
Instead of using NCore or the national NO2, SO2, and PM monitoring networks, national
scale N and S deposition have relied on monitoring networks specifically designed for
estimating deposition. Table 2-3 lists monitoring networks that have been used for recent
estimates of atmospheric N and S deposition for the National Acid Deposition Program
(Schwedc and Lear. 2014a).
2-29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
Table 2-3 Summary of monitoring networks used by Schwede and Lear
(2014a).
Network
Chemical Species Period of Record
Website
CASTNET
Concentration: HNO3, 2000-2012 http://epa.gov/castnet/iavaweb/index.html
SO2, pSCM2"3, pN03~a,
pNH4+a
AMoN
Concentration: NH3
2008-2012 http://nadp.sws.uiuc.edu/AMon/
SEARCH
Concentration: HNO3, 2005-2011
SO2, NH3
NTN
Wet deposition:
SO42", NO3-, NH4+
2000-2012 http://nadp.sws.uiuc.edu/NTN/
apS042- is particulate sulfate concentration, pN03" is particulate nitrate concentration, pNH4+ is particulate ammonium concentration.
Note: summary of data from monitoring networks used in the methodology.
Source: Schwede and Lear (20143).
Wet deposition is estimated as the product of pollutant concentration in precipitation and
precipitation depth (e.g., in rain or snow). Concentration in precipitation is currently
measured as a weekly average by the National Atmospheric Deposition Program/National
Trends Network (NADP/NTN) across a national network of 250 sites using a standard
precipitation collector described in the 2008 ISA (U.S. EPA. 2008a). The NADP
precipitation network was initiated in 1978 to collect data on amounts, trends, and
distributions of acids, nutrients, and cations in precipitation. It expanded to meet the
needs of the National Acid Precipitation Assessment Program established in 1981 to
understand causes and effects of acid precipitation. The NTN is the only network
providing a long-term record of precipitation chemistry across the U.S. Sites are mainly
located away from urban areas and pollution sources. An automated collector ensures that
the sample is exposed only during precipitation (wet-only sampling). Species measured
are free acidity (H+ as pH), conductance, calcium (Ca2+), magnesium (Mg2+), sodium
(Na+), potassium (K+), sulfate (SO42 ). nitrate (NO, ). chloride (Cr), and ammonium
(NH4+). Relatively high confidence has been assigned to wet deposition estimates because
of established capabilities for measuring relevant chemical components in precipitation
samples (U.S. EPA. 2011a). The Atmospheric Integrated Research Monitoring Network
(AIRMoN) started in 1992 and measures the same species as the NTN, but on a daily
rather than weekly basis.
In contrast, direct measurements of dry deposition flux are rare and difficult, and dry
deposition fluxes of gases and particles are estimated from concentration measurements
by an inferential technique described in the 2008 ISA (U.S. EPA. 2008a). In the
June 2018
2-30
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
inferential model approach, the deposition of a pollutant is accomplished by introducing a
resistance component to account for the individual chemical and biological processes that
control pollutant adsorption and capture at natural surfaces (Hicks et al.. 1987).
Concentrations are measured in the Clean Air Status and Trends Network (CASTNET),
which was established under the 1991 Clean Air Act Amendments to assess trends in
acidic deposition. CASTNET is a long-term environmental monitoring network with
95 sites located throughout the U.S. and Canada, managed and operated by the U.S. EPA
in cooperation with other federal, state, and local partners (www.epa.gov/castnet)
including six Native American tribes. CASTNET is the only network in the U.S. that
provides a consistent, long-term data record of acidic dry deposition fluxes. It
complements the NTN, and nearly all CASTNET sites are collocated with or near an
NTN site. Together, these two monitoring programs are designed to provide data
necessary to estimate long-term temporal and spatial trends in total deposition (dry and
wet) as well as ecosystem health. Species measured in CASTNET include: O3, SO2,
HNO3 in the gas phase and SO42 . NO3 . NH4+, Ca2+, Mg2+, K+, Na+, and CI" in particles.
While CASTNET data are more useful for estimating dry deposition than data from FRM
networks, monitors are generally sparse and deposition is only determined for discrete
locations. Also, not all of the species that contribute to total sulfur and nitrogen
deposition are measured in CASTNET (Schwedc etal.. 2011). Despite these
disadvantages, CASTNET data still be very useful if used in combination with modeled
data (Schwedc etal.. 2011). NH3 is not measured in CASTNET, but the National
Atmospheric Deposition Program (NADP) deployed a separate NH3 monitoring network
(AMoN) using Radiello® passive samplers starting in the fall of 2007 at 16 sites;
currently there are more than 60 active AMoN sites, two-thirds of which are located at
CASTNET sites.
A limitation of dry deposition derived from CASTNET and other dry deposition
networks is that results cannot be spatially interpolated because of the complexity of the
deposition field (Schwede and Lear. 2014a; Baumgardner et al.. 2002). Combined with
the sparse coverage of the network, this complexity restricts the capability of routine
monitoring networks to provide data on dry deposition. To some extent, this limitation
can be addressed by considering data from other networks.
The remaining network in Table 2-3 is the Southeastern Aerosol Research and
Characterization network (SEARCH), which is a highly instrumented network of four
urban and four rural stations in Alabama, Florida, Georgia, and Mississippi (Hansen et
al.. 2003). The four rural SEARCH sites have been used for dry deposition estimates.
SEARCH began as a public-private collaboration in early 1998 and has continued
2-31
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
operation with several objectives, including understanding processes governing PM2 5 and
copollutants emissions, transport, and deposition in the southeastern U.S.
One additional network that has been identified as potentially suitable for use in future
deposition estimates (Schw edc et al.. 2011) is the Interagency Monitoring of Protected
Visual Environments (IMPROVE) network. The IMPROVE network consists of more
than 100 monitoring sites in national parks and other remote locations and is primarily
focused on visibility impairment, but has also provided a reliable, long-term record of
particulate mass and species components. Several other monitoring networks are operated
either by the U.S. EPA or jointly with other federal agencies; species measured and other
details for networks making measurements relevant for deposition are shown in Naess
(2016). Even if concentration data from other networks are combined with CASTNET
data, large areas of the U.S. are still relatively far away from, or in a different
environment than, the nearest monitor.
Another deficiency of both the NTN and CASTNET is that not all species that contribute
to total sulfur and nitrogen deposition are measured. Reliable measurements of NOy and
NO2 concentrations, especially at the low concentrations observed in many areas far from
sources, are crucial for evaluating the performance of three-dimensional, chemical
transport models of oxidant and acid production in the atmosphere. To meet this need,
NOy monitors have been installed at six sites in CASTNET as part of the NCore
program. At most sites, however, NO2 is not currently monitored in CASTNET. The
same is true for HNO2 and peroxyacyl nitrates, which can also contribute significantly to
total gas-phase reactive nitrogen. These species can be important contributors to N
deposition locally, especially near populated areas. Neither the NTN nor CASTNET
monitor reduced organic nitrogen compounds, which can also contribute significantly to
N deposition (see Appendix 2.3.4V The sparse geographic coverage and lack of
measurements for key species in these networks along with the awareness of modeling
uncertainties led to the initiation of the Total Deposition Science Committee (TDEP)
(NADP. 2016) to develop hybrid approaches to improve estimates of atmospheric
deposition. The TDEP approach is described in Appendix 2.6.
2.4.2. NO2, NOx, and NOy
2.4.2.1. Network Monitoring
As described in Appendix 2.4.1. a nationwide monitoring network is in place for routine
monitoring of NO2, and NOy is measured in the nationwide NCore network. NO2 is
routinely measured using the FRM chemiluminescence method based on the catalytic
June 2018
2-32
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
reduction of NO2 to NO, followed by reaction NO with O3. However, the reduction of
NO2 to NO on the MoOx catalyst substrate also reduces other oxidized nitrogen
compounds (i.e., NOz compounds shown in the outer box of Figure 2-3) to NO. This
interference by NOz compounds has long been recognized following Winer etal. (1974)
who found NO was also produced by catalytic reduction of HNO3, PAN, and organic
nitrates using this method. As a result of their experiments, Winer et al. (1974) concluded
that, "the NOx mode of commercial chemiluminescent analyzers must be viewed to a
good approximation as measuring total gas-phase 'oxides of nitrogen,' not simply the
sum of NO and NO2." Numerous later studies, as noted in the ISA for Oxides of Nitrogen
(U.S. EPA. 2016f). have confirmed this conclusion. Further details were also described in
the 2008 ISA (U.S. EPA. 2008a).
Commercially available NOx monitors have been converted to NOy monitors by moving
the molybdenum oxide catalyst substrate to interface directly with the sample inlet to
improve the efficiency of reduction of NOz compounds susceptible to loss on inlet
surfaces. NOx concentrations cannot be considered as a universal surrogate for NOy.
However, near sources of fresh combustion emissions, such as highways, most of the
NOy is present as NOx. To the extent that all the major oxidized nitrogen species can be
reduced quantitatively to NO, measurements of NOy concentrations should be more
reliable than those for NOx concentrations, particularly at typical ambient levels of NO2.
Exceptions might apply in locations near NOx sources, where NOx measurements are
likely to be less biased and confidence in measurement accuracy increases.
2.4.2.2. Remote Sensing
Satellite-based methods have also been used to measure NO2. Remote sensing by
satellites is especially useful in areas where surface monitors are sparse. Retrieving NO2
column abundances from satellite data typically involves three steps: (1) determining the
total NO2 integrated line-of-sight (slant) abundance by spectral fitting of measurements
of backscattered solar radiation, (2) removing the stratospheric contribution by using data
from remote regions where the tropospheric column abundance is small, and (3) applying
an air mass factor to convert tropospheric slant columns into vertical columns. The
retrieval uncertainty is largely determined by steps 1 and 2 over remote regions where
there is little tropospheric NO2, and by step 3, over regions of elevated tropospheric NO2
(Boersma et al.. 2004; Martin et al.. 2002). Satellite retrievals are largely limited to cloud
fractions <20%. A hybrid approach using data for NO2 tropospheric column abundances
obtained by the Ozone Monitoring Instrument (OMI) on the Aura satellite coupled with
results from the GEOS-Chem, global-scale, three-dimensional, chemistry-transport model
has been developed by Lamsal et al. (2008) with updates by Lamsal et al. (2010). In this
June 2018
2-33
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
approach, the surface mixing ratio divided by the column abundances calculated by the
GEOS-Chem model are used as the scaling factors to derive surface mixing ratios from
satellite-measured column abundances. This method provides estimates of surface NO2
concentrations that are especially useful in data-sparse regions. The algorithm used to
derive the tropospheric columns of NO2 is given in Bucsela et al. (2013). Note that this
algorithm was recently shown to produce NO2 column abundances that are too high by
-20% (Marchcnko et al.. 2015).
2.4.2.3. Research and Nonroutine Methods
Alternatively, multiple methods for observing components of NOy have been developed
and evaluated in some detail. As a result of these methods, as applied in the field and the
laboratory, knowledge of the chemistry of odd-N species has evolved rapidly. Recent
evaluations of methods can be found in Arnold et al. (2007) for HNO3; Wooldridge etal.
(2010) for speciated PANs; and Pinto et al. (2014) for HONO. However, it is worth
reiterating that the direct measurements of NO are still the most reliable of all.
2.4.3. Ammonia
2.4.3.1. Network Monitoring
The recently implemented AMoN for monitoring ammonia was described in
Appendix 2.4.1. The passive sampling method relies on diffusion across a membrane
onto an absorbing substrate, which for NH3 is H3PO3. The sampling period in AMoN is
2 weeks. Puchalski et al. (2011) compared the results from three passive samplers with
annular denuder systems (taken to be the reference method). The median relative
percentage difference between the Radiello passive samplers and the denuder systems
was -37% and the coefficient of variation among triplicate Radiello samplers was 10%.
Puchalski et al. (2015) further compared 2-week samples collected at five sites over the
course of a year by Radiello passive samplers with collocated annular denuder systems
(ADS) with different configurations. The mean relative percentage difference between
the ADS and AMoN samplers was -9% to be compared to a precision of 5% for both the
ADS and AMoN samplers.
June 2018
2-34
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
2.4.3.2.
Remote Sensing
In addition to these in situ techniques, remote sensing techniques have also been used to
measure NH3. The Tropospheric Emission Spectrometer (TES) on the Aura satellite
rShephard et al. (2011). and references therein; (Beer et al.. 2008)1 and the Infrared
Atmospheric Sounding Interferometer (IASI) on the MetOp-A satellite (Van Damme et
al.. 2014) measured spectral features in the v2 vibrational band centered at around
950/cm (the so-called atmospheric window in the infrared). Operating specifications for
TES (spectral resolution, 0.06/cm; footprint 5.3 x 8.3 km2; 0.15-0.20 K noise) are
generally better than for IASI (spectral resolution, 0.50/cm; footprint 12 x 12 km2;
0.15-0.20 K noise). Although TES has higher spectral resolution, it has less dense spatial
coverage. Unlike satellite detection of atmospheric molecules by backscattered solar
radiation (e.g., NO2 and SO2), NH3 is detected in the thermal infrared spectral range, so
data for both day and night can be obtained (satellite overpasses at the Equator at
approximately 1:30 a.m. and 1:30 p.m. for TES and 9:30 a.m. and 9:30 p.m. for IASI).
The sensitivity of the IR sounding technique for NH3 increases with the thermal contrast
between the surface and the temperature of the air in the lower troposphere, and thus the
daytime crossing allows for increased detectability of NH3 (Clarisse et al.. 2010). NH3 is
confined largely to the planetary boundary layer (PBL), with much lower concentrations
aloft in the free troposphere. TES retrievals are most sensitive to NH3 at atmospheric
pressures between 700 and 900 hPa. The TES level of detectability for NH3 is given by a
profile with a peak concentration of 1 ppbv, or equivalently a constant mixing ratio of
0.4 ppbv distributed over the pressure range of maximum sensitivity, provided there is
substantial thermal contrast.
Pinder etal. (2011) found that TES retrievals of NH3 in the PBL captured the spatial and
seasonal variability of NH3 over eastern North Carolina measured by surface
observations. Similarly, Sun et al. (2015a') found that column abundances measured by
TES over the San Joaquin Valley agreed with those measured by upward looking
instruments at the surface to within 2% and to within 6% for aircraft measurements. TES
columns were also shown to be reasonably well correlated (R2 = 0.67) with median NH3
measured at the surface by quantum cascade laser, thereby demonstrating the ability of
the satellite-based measurements to capture spatial variability in NH3 between individual
pixels.
Alvarado et al. (2011) derived emissions factors for NH3 in biomass burning plumes over
Canada using data from TES. Using data from IASI R'Honi et al. (2013) found total
column abundances for NH3 in the plumes of Russian wildfires during the summer of
2010 that were two orders of magnitude larger than background values.
June 2018
2-35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
2.4.3.3.
Research and Nonroutine Methods
Ambient instruments with much higher time resolution were compared by Schwab et al.
(2007) and von Bobrutzki et al. (2010). Schwab et al. (2007) conducted a
laboratory-based intercomparison of ambient NH3 instruments with seven instruments
using six methods sampling from a common manifold, including tunable diode laser
(TDLAS) absorption spectrometer, wet scrubbing long-path absorption photometer
(LOPAP), wet effusive diffusion denuder (WEDD), ion mobility spectrometer (IMS),
Nitrolux laser acousto-optical absorption analyzer, and a modified CL analyzer. Schwab
et al. (2007) reported that all instruments agreed to within -25% of the expected
calibration value, with the exception of the CL analyzer which suffered from problems
related to its MoOx conversion of NOz to NO.
von Bobrutzki et al. (2010) conducted a field intercomparison of ambient NH3
measurements with 11 instruments using 8 methods including: 3 wet techniques (annular
rotating batch denuders, 1 with offline analysis and 2 with online analysis [AMANDA,
AiRRmonia]), 2 Quantum Cascade Laser Absorption Spectrometers (c-QCLAS,
DUAL-QCLAS), 2 photo-acoustic spectrometers, a cavity ring down spectrometer, a
chemical ionization mass spectrometer, an ion mobility spectrometer, and an open-path
Fourier transform infrared spectrometer. This study was unique in that the surrounding
field was fertilized with urea halfway through the campaign to increase average
concentrations of NH3 from 10 to 100 ppb. Overall, If was >0.84 with respect to the
ensemble mean for all instruments over the entire range of concentrations (<120 ppb),
with slopes ranging from 0.67 to 1.13. Higher variability was found at lower
concentrations (<10 ppb) with R2 > 0.52 and slopes ranging from 0.42 to 1.15. Perhaps
the most consistent agreement between two instruments was found for the c-QCLAS and
AiRRmonia (R2 = 0.91, slope = 0.86, intercept = 0.84 ppb for NH3 <10 ppb; and
R2 = 0.91, slope = 0.83, intercept = 0.34 ppb over the entire range ofNFL concentrations).
Nitrolux-100 denuders were used both in the intercomparison study of von Bobrutzki et
al. (2010) and in the one by Puchalski et al. (2011). Compared to the ensemble mean, the
slope was 0.97, the intercept was 1.86 ppb, and R2 = 0.98 over the entire concentration
range. As noted by von Bobrutzki etal. (2010). comparisons of this sort only show
relative performance of the instruments and not a functional relationship to a standard.
June 2018
2-36
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
2.4.4.
Sulfur Dioxide
2.4.4.1. Network Monitoring
In the nationwide monitoring network for SO2, SO2 is routinely measured by pulsed UV
fluorescence. This technique is a Federal Equivalence Method (FEM). The method's
principles and potential inteferences have been described in detail in the 2008 ISA (U.S.
EPA. 2008a'). However, measurements using this method are not as widely used as
measurements from the CASTNET. In the CASTNET, SO2 is collected by capturing it on
filters and measured as sulfate following procedures described in Appendix 2.4.5.
2.4.4.2. Remote Sensing
In addition to the above in situ methods, satellite-based measurements have also been
used to measure tropospheric SO2 and to infer surface SO2 concentrations with the aid of
the GEOS-Chem chemistry-transport model (Now lan et al.. 2014; Lee etal.. 2011).
Tropospheric column abundances of SO2 are obtained by the Ozone Monitoring
Instrument (OMI) on the Aura satellite or the Scanning Imaging Absorption Spectrometer
for Atmospheric Cartography (SCIAMACHY) on Envisat and are combined with results
from the GEOS-Chem, global-scale, three-dimensional, chemistry-transport model to
derive surface concentrations of SO2 (as they are for NO2). Lee et al. (2011) associated
annual mean surface mixing ratios of SO2 derived from the hybrid satellite/model
technique with ambient measurements of SO2, (R2 = 0.66 and 0.74, slope = 0.70 and 0.93,
n = 121 and 115, for OMI and SCIAMACHY, respectively).
The algorithms used to derive vertically integrated SO2 abundances in the troposphere
undergo continuing refinement. For example, Thevs et al. (2015) applied an algorithm
based on differential optical absorption spectroscopy (Piatt and Stutz. 2008) combined
with a radiative transfer model. Li et al. (2013) developed an algorithm based on
principal components analysis, which has replaced the earlier standard algorithm
developed by Krotkov et al. (2008). The methods applied by Thevs et al. (2015) and Li et
al. (2013) are designed to be operational for retrieving SO2 in the PBL with an estimated
detection limit of ~0.5 Dobson units (1 DU = 2.69 x 1016 molec/cm2 corresponding to a
concentration of ~3 ppb if SO2 is well mixed in a 2-km-deep mixed layer), or about half
that in the older standard method. Table 2-4 summarizes sources of uncertainty for
individual OMI measurements of NO2 and SO2.
June 2018
2-37
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
Table 2-4 Sources of uncertainty for individual Ozone Monitoring Instrument
measurements in the studv of Nowlan et al. (2014).
Source
NO2
SO2
OMI fitting error
~1015 molec/cm2
~1016 molec/cm2
Air mass factor
20%
15-45%
Stratospheric correction
2 x 1014 molec/cm2
N/A
SO2 offset correction
N/A 2
x 1014 molec/cm2
Profile shape
30%
10%
molec = molecules; N/A = not applicable; N02 = nitrogen dioxide; OMI = Ozone Monitoring Instrument; S02 = sulfur dioxide.
Source:Nowlan et al. (2014).
The errors in the column measurements result mainly from uncertainties in the vertical
profiles of NO2 and SO2, cloud fraction, cloud pressure, surface reflectivity, and particles
used in the calculation of air mass factor. A correction is required to account for NO2 in
the stratosphere (produced from N2O oxidation and cosmic ray interactions dissociating
with N2). The SO2 offset correction refers to a global background correction arising from
issues in spectral fitting, such as spectral correlations with O3 and stray light within the
instrument.
2.4.5. Filter-Based Concentration Measurements
As described in Appendix 2.4.1. most measurements used for estimating deposition are
from CASTNET, rather than from monitoring networks based on FRM and FEM
methods. The CASTNET filter pack is shown in Figure 2-6. Particulate matter is
collected on the open-face Teflon filter, extracted in deionized water, and analyzed by ion
chromatography (IC) for sulfate, nitrate, ammonium, and other species identified in
Figure 2-6. In the CASTNET filter pack, gases are collected downstream of the
particulate species, with nitric acid collected on nylon filters and analyzed as NO3 by ion
chromatography, and SO2 on carbonate impregnated cellulose filters and analyzed as
S042 by ion chromatography. Extensive intercomparisons of CASTNET methods with
other measurement methods were described in the 2008 ISA (U.S. EPA. 2008a).
June 2018
2-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
Figure 1-4 Three-Stage Filter Pack
Cellulose (2)
Nylon
Teflon®
Gaseous
Gaseous
Particulate
•so3
• hno3 ¦ so2
• so, • no3 • nh; • K*


• Ca2* • Mg:T • Na* • Cr
Cellulose Nyton Teflon*
Quick Disconnect	Fillers	Filter	Filter	Shipping Cap
Ca2t = calcium ion; CI = chloride; HN03 = nitric acid; K+ = potassium ion; Mg2+ = magnesium ion; Na+ = sodium ion:
NH4+ = ammonium; N03" = nitrate; S02 = sulfur dioxide; S042- = sulfate.
Air flows from right to left.
Source: MACTEC (2010).
Figure 2-6 Clean Air Status and Trends Network filter pack
As can be seen in Figure 2-6. SO2 is measured by the CASTNET filter pack by IC
analysis of extracts from the cellulose filters. Because the nylon filter adsorbs some of the
SO2 (Sickles et al.. 1999; Sickles and Hodson. 19991 SO42 is also measured on nylon
and added to the SO2 (expressed as S042 ) collected on the backup cellulose-fiber filters.
Uncertainties in CASTNET data are reported quarterly in a quality assurance report (U.S.
EPA. 2016c). Precision is determined as the absolute value of quarterly or annually
aggregated relative percentage difference for duplicate sample pairs collected with
collocated samplers at two sites. Data quality objectives for ammonium, nitrate, and
sulfate are within 20%, but reported precision for 2016 was well under this target, 2-5%
for sulfate, 5—13% for nitrate, and 2-6% for ammonium, where the range reflects that
there are two sites. Analytical accuracy was reported within 2% based on spiked
calibration verification samples. Further detail on uncertainty and data quality can be
found in (U.S. EPA. 2016c). Additional unknown uncertainty is associated with
volatilization of NH4 from collected PM. Substantial loss during sampling can occur
because collected NH4NO3 in PM is in a temperature dependent equilibrium with NH3
and HNO3 (see Appendix 2.3.3). leading to volatilization of both species after PM
June 2018
2-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
collection. Loss of NH4+ during sampling was thoroughly reviewed in the 2008 ISA (U.S.
EPA. 2008a).
Results of an intercomparison of weekly average SO2 data (ppbv) collected by the
CASTNET filter pack and trace level SO2 monitors during all of 2014 at Bondville, IL
and Beltsville, MD are shown in Figure 2-7 (AMEC Environment & Infrastructure.
2015).
In addition to CASTNET, pSCV", pNOs are monitored in the Chemical Speciation
Network (CSN), and the Interagency Monitoring of Protected Visual Environments
(IMPROVE) network. Sampling and measurement methods for these networks were
described in detail by Solomon et al. (2014). In the CSN network, pNH4+ is also
measured, but as for CASTNet, it may be subject to volatilization error (volatilization of
N03 is corrected).
2-40
DRAFT: Do Not Cite or Quote

-------
Trace UV Fluorescence S02
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
Trace UV Fluorescence SCI;
Filter Pack vs Trace U V Fluorescence S02
Beltsville, MD
±0.05
Filter Pack vs Trace UV Fluorescence S02
Bondville, IL
±0.07
S02 = sulfur dioxide; UV = ultraviolet.
Results for Beltsville, MD are shown in the upper panel and results for Bondville, IL are shown in the lower panel.
Figure 2-7 Comparison between weekly average measurements of sulfur
dioxide using the Clean Air Status and Trends Network filter pack
and the trace ultraviolet pulsed fluorescence monitor in 2014.
June 2018
2-41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
2.4.6. Deposition Measurements
Wet deposition is estimated as the product of pollutant concentration in precipitation and
precipitation depth (e.g., in rain or snow; see Appendix 2.4.1). Concentration in
precipitation is currently measured by the National Atmospheric Deposition
Program/National Trends Network (NADP/NTN) across a national network of sites using
a standard precipitation collector described in the 2008 ISA (U.S. EPA. 2008a').
Relatively high confidence has been assigned to wet deposition estimates because of
established capabilities for measuring relevant chemical components in precipitation
samples (U.S. EPA. 201 la). Measurement precision determined as the average absolute
percentage differences of replicate samples in the 2016 annual quality assurance report
was 1% or less for sulfate, nitrate, and ammonium, and absolute percentage difference
was no greater than 5% for sulfate and ammonium or 7% for nitrate for any single sample
pair (U.S. EPA. 2016c). Bias determined from internal blind samples was 2% or less for
sulfate and nitrate, and 6% or less for ammonium (U.S. EPA. 2016c).
In contrast, direct measurements of dry deposition flux are rare and difficult. Methods for
estimating dry deposition from field measurements fall into two major categories: surface
analysis methods, which include all types of estimates of contaminant accumulation on
surfaces of interest, and atmospheric deposition rate methods, which use measurements
of contaminant concentrations in the atmosphere and micrometeorological measurements
of atmospheric turbulence (U.S. EPA. 2008a). Emphasis here is placed on the latter class
of methods, which are more widely applicable because the accumulation methods are
subject to limitations such as the site specificity of the measurements and the restriction
to elements that are largely conserved within the vegetative system. Dry deposition
estimates using atmospheric deposition rate methods are based on field measurements of
a species or particle concentration gradient along with a measurement or estimate of its
turbulent diffusivity under the field conditions of the measurement (Mvles et al.. 2012;
Businger. 1986). Examples include eddy covariance and aerodynamic gradient
techniques (U.S. EPA. 2008a). In the eddy covariance method, flux is calculated from the
covariance between fluctuations in wind velocity and concentration. Historically,
empirical estimates of deposition for wind tunnel and field conditions often have not
agreed well with theoretical predictions, probably because transport phenomena and
turbulence structure near surfaces are not well characterized (U.S. EPA. 2004). However,
improvement of dry deposition measurements is an active research area, both to reduce
measurement uncertainties and to improve modeling capabilities by better understanding
deposition processes and their parameterization in chemical transport models.
Progress in improving measurement capabilities has been triggered by the development
of continuous air sampling measurement techniques with higher sensitivity and temporal
June 2018
2-42
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
resolution, with the objective of improving costs and measurement quality using
atmospheric deposition rate measurement methods. Concluding that the main reason for
the shortage of dry deposition measurements is the expense and complexity of
measurement methods, (Almand-Huntcr et al.. 2015) tested a dynamic flux chamber
using automated, inexpensive multispecies gas flux monitoring system for measurement
of a variety of pollutants, including NOx, that would be needed to contend with the large
spatial and temporal variability in air-surface exchange rates of reactive compounds. A
dynamic chamber system was also developed to allow measurements of NO, NO2, and O3
measurements so that compensation points and deposition velocities (see Appendix 2.5.2)
could be estimated (Breuninger et al.. 2012).
As part of a series of measurements to be tested at several CASTNET sites, Rumsev and
Walker (2016) evaluated the ability of the MARGA 2S (Monitoring for AeRosols and
GAses) system to simultaneously measure fluxes of NH3-HNO3-NH4NO3 using the
aerodynamic gradient method to allow for an assessment of the errors due to the
instability of the particle phase, as well as SO2 and SO42 . to allow for the investigation of
ammonium sulfate neutralization and codeposition between SO2 and NH3. Over a range
of meteorological conditions, median flux uncertainty was found to range from ~31 % for
NH3 to ~ 120% for NH4+. The flux gradient technique was also applied to a forest clearing
as an example of a complex ecosystem with the objectives of improving deposition rate
accuracy and model parameterization for SO2 (Mvles et al.. 2012). Deposition velocities
fluctuated considerably with a mean of 1.00 ± 0.48 cm/s, and the large variation was not
fully captured by estimates from widely used models (Mvlcs et al.. 2012). Uncertainties
in canopy resistance, including stomatal and nonstomatal processes were identified as
probable sources of uncertainty (Mvlcs et al.. 2012).
This approach has been successfully applied to measurements of a wide variety of
species. However, it is less suitable for HNO3 and NH3 because they are subject to
interactions with inlet surfaces of measuring devices. For example, in one study a
correction factor of 1.62 was reported for inlet surface interactions (Breuninger et al..
2012). To address the problem of loss of HNO3, NH3, and other substances that interact
with sampling inlet surfaces during measurement, Roscioli et al. (2016) developed a
method to passivate inlet surfaces and thereby overcome this difficulty by allowing for
more rapid response measurements. Min et al. (2014) determined the net flux of NO and
NO2 (at a frequency of 5 Hz) over a forest with estimated total systematic uncertainties of
<8 and <6%, and random errors of <25 and <21%, respectively.
Measurement methods are well developed for ideal conditions of flat, homogeneous, and
extensive landscapes and for chemical species for which accurate and rapid sensors are
available (U.S. EPA. 2008a). However, the strong dependence of dry deposition on
2-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
surface characteristics, which are highly spatially variable, causes in situ measurements
to be limited in applicability, especially in areas of high topographic relief
[e.g., (Weathers et al.. 2006)1. The lifetime in the boundary layer (~1 km depth) for
rapidly depositing species is ~1 day, implying that distances from source to deposition
are relatively short and that the deposition flux depends strongly on the nature and
strength of nearby sources. Deposition also shows strong temporal variability on
timescales ranging from diurnal to seasonal, as do wind regimes affecting a particular
site. In addition, cost and logistics make the eddy covariance and aerodynamic gradient
techniques impractical for monitoring networks.
Instead, dry deposition fluxes of gases and particles are estimated in CASTNET and by
chemistry-transport models, such as CMAQ, by an inferential technique described in the
2008 ISA (U.S. EPA. 2008a). In the inferential model approach, the deposition of a
pollutant is estimated by introducing a resistance component to account for the individual
chemical and biological processes that control pollutant adsorption and capture at natural
surfaces (Hicks et al.. 1987). Ambient pollutant concentrations of O3, SO42 . NO3 . NH4+,
SO2, and HNO3 are routinely collected at CASTNET dry deposition sites. Deposition
velocities based on local meteorological measurements were calculated using the
Multi-Layer Model (MLM) at U.S. EPA-sponsored CASTNET sites until 2010, when
meteorological measurements were discontinued at all but five U.S. EPA CASTNET
sites. Dry deposition fluxes are still reported at sites with discontinued meteorological
measurements using historical data. A disadvantage to this approach is that relevant
atmospheric species are not routinely measured. For example, NO2 and peroxyacetyl
nitrate, which together consistently contribute 15 to 25% of estimated oxidized nitrogen
dry deposition, are not measured at CASTNET sites. Even for those species that are
routinely measured, network spatial coverage is sparse (U.S. EPA. 2011a).
Satellite-based measurements offer a potential means of greater coverage, but only a
limited number of deposition related applications have been described. (Nowlan et al..
2014) combined satellite data with modeled NO2 and SO2 surfaces and vertically
integrated concentrations and deposition velocities to estimate deposition fluxes. SO2 dry
deposition fluxes compared well with surface network based deposition fluxes.
Uncertainties in depositional flux estimates in this approach result from the combined
uncertainties in the satellite-derived surface concentrations and model-derived deposition
velocities and were estimated to be -30% on average for both NO2 and SO2 over land. In
the absence of routine measurements, dry deposition is often modeled with CMAQ or
other modeling tools using relevant emissions, meteorological, and land use data, rather
than estimated from measured concentration measurements (U.S. EPA. 2011a).
2-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Modeling dry deposition is particularly challenging over varying terrain and under
extremely stable conditions such as those occurring at night. Under optimal conditions
over a relatively small area where dry deposition measurements have been made,
uncertainties on the order of ±30% have been reported and larger uncertainties are likely
when the surface features in the built environment are not well known or when the
surface comprises a patchwork of different surface types, as is common in the eastern
U.S. (U.S. EPA. 2008a). For this reason, dry deposition is routinely estimated from
concentration measurements, usually from CASTNET data (Appendix 2.4.5) using a
hybrid approach based on both measured and modeled data (Appendix 2.6).
2.5. Modeling Chemistry, Transport, and Deposition
In this section, advances in modeling transport and deposition of species relevant to acid
and nutrient deposition are discussed, along with research progress on understanding
underlying transport and deposition processes. The use of chemical transport models
(CTMs) to model deposition was discussed extensively in the 2008 ISA (U.S. EPA.
2008a). Relevant new research and improvements in CTM modeling in general are
described in Appendix 2.5.1. Appendix 2.5.2 discusses environmental processes relevant
to understanding and modeling acid and nutrient deposition. Appendix 2.5.2.1 begins
with an overview of fundamental processes of atmospheric deposition of gases and
particles, along with deposition velocities for some key gas-phase species. It also contains
discussions of research advances in three key processes that serve as major structural
uncertainties (lack of knowledge of the underlying science) in modeling deposition: NOx
canopy processes, which involve both bidirectional gas exchange and NOx chemistry
(Appendix 2.5.2.2); bidirectional exchange of ammonia (Appendix 2.5.2.3); and
transference ratios relating average ambient concentration to deposition flux
(Appendix 2.5.2.4). Appendix 2.5.3 discusses model evaluation and uncertainty,
including comparisons between CTM and network-based wet deposition results.
2.5.1. Advances in Chemistry-Transport Model (CTM) Modeling
To understand the lifetime and fate of the varied forms of atmospheric sulfur and nitrogen
from emission to deposition, it is necessary to account for both atmospheric transport and
chemical transformations. CTMs simulate the relevant atmospheric transport processes
(e.g., horizontal and vertical advection and diffusion), as well as chemistry, aerosol
physics, deposition, and cloud processes. The 2008 ISA (U.S. EPA. 2008a) provided a
detailed description of the CTM models and their application to estimating deposition.
Continental-scale CTMs include CMAQ (Appel et al.. 2017) and CAMx (Koo et al..
June 2018
2-45
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
2015). while GEOS-Chem (Zhang et al.. 2012a) is an example of a global-scale model.
Most major regional-scale, air-related modeling efforts at U.S. EPA use the Community
Multiscale Air Quality (CMAQ) modeling system (Bvun and Schere. 2006; Bvun and
China. 1999). Recent updates to CTM model design, and in particular to CMAQ, are
described here.
A number of complex atmospheric processes influence pollutant behavior between
emission and deposition and must be taken into account to achieve good model
performance. A variety of mechanisms operating over a wide range of spatial and
temporal scales transport heat, water, and pollutants horizontally and vertically through
the atmosphere. These mechanisms range from local-scale circulations (e.g., urban heat
islands) to hemispheric-scale transport by the jet stream. Long-range transport of
pollutants in the lower free troposphere associated with large-scale synoptic systems is
possible because flows are largely uncoupled from surface friction. Flows in the upper
planetary boundary layer (PBL), especially during the day, might not be as effective for
transporting pollutants over long distances because air can be mixed down to the surface
by turbulence. If these pollutants react with surface material or are taken up by
vegetation, they can be removed within a relatively short distance from their sources.
In addition to wind velocity, the distance scale for transport of a pollutant that is
relatively stable in the troposphere with respect to gas-phase reactions (i.e., chemical
lifetime > a few days) depends strongly on the pollutant's interactions with solid and
liquid surfaces and subsequent chemical transformations. Because NO and NO2 are only
slightly soluble, they can be transported over longer distances in the gas phase than can
more soluble species like HNO3 (and its anhydride, N2O5) and NH3 that are depleted by
deposition to moist surfaces or taken up more readily on aqueous surfaces of particles or
on cloud drops. For example, measurements of the ratio of NH3 to CH4 in the San
Joaquin Valley indicate substantial loss of NH3 to the surface within a few km of sources
of these gases (Miller et al.. 2015). consistent with an NH3 lifetime of minutes to a few
hours in this environment. On the other hand, a combination of models, remote sensing,
and in situ measurements over the eastern U.S. indicate an atmospheric lifetime for SO2
of 19 ± 7 hours in summer increasing to 58 ± 20 hours in winter (Lee et al.. 2011). which
indicates the potential for much longer-range transport of SO2.
Numerous advances in atmospheric science have been codified in CTMs, including
gas-phase oxidant chemistry relevant for the formation of aerosol precursors and dry
deposition by gravitational settling (Nolte et al.. 2015). improved representation of
meteorological processes in CTMs and interactions with aerosols (Tuccella et al.. 2015).
and improved algorithms for understanding the influence of weather on emissions of
ammonia from agricultural lands (Flechard et al.. 2013). Over the U.S. and Europe,
2-46
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
substantial reductions in sulfur dioxide and nitrogen oxides have created an opportunity
to compare the model results with the trends in ambient observations (Banzhaf et al..
2015; Xing et al.. 2015; Civerolo et al.. 2010). Studies have shown that CMAQ is skilled
at capturing the seasonal and long-term changes in sulfate PM2 5 and nitrate PM2 5 as well
as total PM2 5 mass; however, the model performs less well for seasonal variability in
nitrate PM2 5, owing to uncertainties in ammonia emission trends (Banzhaf et al.. 2015;
Xing et al.. 2015).
2.5.2.
2.5.2.1.
Considerable advances in both our understanding of atmospheric deposition and
modeling approaches to characterize it have taken place recently. Deposition is a
complicated process influenced by numerous atmospheric and deposition surface
properties, as well as chemical reactions and other processes that take place within
canopies of vegetation.
In Figure 2-8. Moller (2014) illustrates the pathways that transfer gaseous and particulate
pollutants from the atmosphere to the surface by deposition. Wet deposition occurs when
particulate and gaseous species are removed by cloud drops or by falling precipitation
(washout). Dry deposition occurs when they are removed without precipitation by
processes like turbulence and gravitational settling. In mountainous areas, a third
important type of deposition occurs, referred to as occult deposition (Pollard et al..
1983). which is not shown in Figure 2-8. and results from the impaction of droplets in
fogs or clouds on vegetation. Wet deposition is determined as the simple product of
concentration in precipitation and precipitation rate. Receptor (i.e., vegetation) surface
properties have little effect on wet deposition. Dry deposition is more difficult to
determine. It can be described as a flux Fd, the mass of pollutant deposited per unit area
of the Earth's surfaces where it deposits, or a deposition velocity that relates the dry
deposition flux Fd to a pollutant's ambient concentration:
Fd = vdC
Equation 2-12
where C is the pollutant's concentration in mass in per unit volume, and Vd is the
deposition velocity, which relates a pollutant's deposition flux to its ambient
concentration.
Modeling Deposition
Wet and Dry Deposition of Gases and Particles
June 2018
2-47
DRAFT: Do Not Cite or Quote

-------
1	Dry deposition of gases depends on leaf area, surface resistance to gas uptake,
2	interactions with biota through both stomatal and plant surface pathways, and
3	atmospheric reactivity, which can vary among depositing gases. Measurements of
4	average dry deposition velocities for gases over land surfaces are shown in Table 2-5. and
5	an indication of the seasonal variability of deposition velocities over different land
6	surface types can be seen in Table 2-6. which shows the variability in Vd for SO2.
7	Deposition velocities of other species are also expected to be spatially and temporally
8	variable.
Source: Moller (2014).
Figure 2-8 Schematic diagram showing mechanisms for transferring
pollutants from the atmosphere to the surface.
June 2018
2-48
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Table 2-5 Average dry deposition velocities (cm/s) for several gases over land
surfaces.
Substance
SO2
NO
NO2
HNOs
Os
H2O2
CO
NHs
Vd
0.8
<0.02
0.02
3
0.6
2
<0.02
1
CO = carbon monoxide; H202 = hydrogen peroxide; HN03 = nitric acid; NH3 = ammonia; NO = nitric oxide; N02 = nitrogen dioxide;
03 = ozone; s = second; S02 = sulfur dioxide; Vd = deposition velocities.
Source: Moller (2014).
Table 2-6 Deposition velocity (cm/s) for sulfur dioxide averaged over different
land use types for summer and winter.

Farmland
Grassland
Dec. Forest
Con.
Forest
Urban
Water

Su
Wi
Su
Wi
Su
Wi
Su
Wi
Su
Wi
Su
Wi
Wet
1.0
1.0
1.0
1.0
3.0
1.5
2.0
2.0
1.0
1.0
0.5
0.5
Dry
0.7
0.5
0.6
0.4
1.5
0.5
0.7
0.5
0.1
0.1
0.5
0.5
Snow
-
0.
-
0.1
-
0.2
-
0.2
-
0.1
-
0.1
Con. = coniferous; Dec. = deciduous; Su = summer; Wi = winter.
Source: Moller (2014).
A wide range of deposition velocities is observed among different atmospheric gas-phase
species. HNO3 is an example of gas with straightforward deposition behavior. It sticks
easily to vegetative surfaces, i.e., there is a negligible surface resistance to HNO3 uptake
by vegetation, and its deposition rates are independent of leaf area or stomatal
conductance, implying that deposition occurs to branches, soil, and the leaf cuticle as
well as leaf surfaces. The HNO3 Vd typically exceeds 1 cm/s and exhibits a diel pattern
controlled by turbulence characteristics of midday maxima and lower values at night in
the more stable boundary layer (U.S. EPA. 2008a. 2004). In contrast, NO2 interaction
with vegetation is more complicated. The uptake rate by foliage is related to stomatal
conductance and is more variable. It may also be associated with concentrations of
reactive species such as ascorbate in the plant tissue that react with NO2. At very low
NO2 concentrations, emission from foliage is observed. Internal NO2 appears to derive
from plant N metabolism, and there is evidence for a compensation point, typically near
June 2018
2-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
1 ppb, at which uptake and emission rates are equal and net flux is zero (U.S. EPA.
Dry deposition of PM is influenced by a number of variables, including particle diameter,
atmospheric stability, deposition surface roughness, and the shape, stickiness, roughness,
and cross-sectional area of leaves. Greater roughness and leaf shape complexity increase
deposition. The diversity of particle sizes, atmospheric conditions, and surface
characteristics makes it difficult to estimate dry deposition (U.S. EPA. 2008a. 2004). The
appreciable effects of particle size, local micrometeorological conditions and surface
characteristics on deposition velocity can be seen in Figure 2-9. The key to measurements
of Vd over surfaces covered by low vegetation is given in the left column and over forest
in the right column. These measurements are compared to six model formulations which
are shown as lines in the center column.
For particles >10 |im. Fj varies between 0.5 and 1.1 cm/s, and a minimum particle Ki of
0.03 cm/s exists for particles in the size range 0.1 to 1.0 |im. while deposition of particles
from the atmosphere to a forest canopy has been estimated as 2 to 16 times greater than
deposition in adjacent open terrain like grasslands or other low vegetation (U.S. EPA.
2004).
2-50
DRAFT: Do Not Cite or Quote

-------
A
~
O
o
grass (Chamberlain 67)
grass (WT) "
sticky grass (WT) "
grass (WT) (Clough 75)
moorland (Gallagher 88)
moorland (Nemitz02)
semi arid savannah (Lamaud 94)
Davidson 82 u»=45 cm.s"''
Legg 79-Aylor 82 "
Wiman 85 "
Slinn 82 "
Zhang 01 "
•	Spruce (Beswick 91)
®	Spruce (Bleyl 01)
O	Pine (Lorenz 89)
~	Pine (Buzorius 00)
O	Pine (Gaman 04)
v	Pine (Lamaud 94)
o	Pine (Gronholm 07)
A	Fir (Gallagher 97)
01	0.1	1	10	100
dp (nm)
01	0.1	1	10
dp (nm)
0.001
100
0.001
0.
i—i—i—			1—		
: Grass
dp = aerodynamic diameter of particle; Vdc = deposition velocity; Ws = Stokes settling velocity.
Notes: (Gronholm et al.. 2007: Gaman et al.. 2004: Nemitz et al.. 2002: Zhang et al.. 2001: Buzorius et al.. 1998: Gallagher et al..
1997: Lamaud et al.. 1994a: Lamaud et al.. 1994b: Beswick et al.. 1991: Lorenz and Murphy. 1989: Gallagher et al.. 1988: Wiman
and Agren. 1985: Avlor. 1982: Davidson et al.. 1982: Legg and Powell. 1979: Clough. 1975: Chamberlain. 1967).
Closed symbols correspond to wet or sticky surfaces or liquid particles; open symbols to dry surfaces or solid particles.
Source: Petroff et al. (2008).
Figure 2-9 Modeled and measured deposition velocities over grass (left
figure) and coniferous forest canopies (right figure) for particles
of density 1 am/cm3 depositing under similar friction velocity (u*)
(35 < u* < 56 cm/s).
2.5.2.2. NOx Canopy Processes
1	There are a number of ways that landscape characteristics influence the deposition
2	process (U.S. EPA. 2008a. 2004). In terrain containing extensive vegetative canopies,
3	any material deposited via precipitation to the upper stratum of foliage is likely to be
4	intercepted by several foliar surfaces before reaching the soil. Ki is usually greater for a
5	forest than for a nonforested area and greater for a field than for a water surface. The
6	upwind leading edges of forests, hedgerows, and individual plants are primary sites of
June 2018
2-51
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
coarse particle deposition, and upper canopy foliage tends to receive maximum exposure
to coarse and fine particles, but foliage within the canopy tends to receive primarily fine
particles. (U.S. EPA. 2008a. 2004). Several Nr species are deposited to vegetation, among
them HNO3, NO2, PAN (and other RONO2), and NH3.
Wet, dry, and occult deposition all contribute N and S species to the forest canopy in
varying proportions. Deposited species can react on surfaces within the canopy, be taken
up by vegetation through stomata, be resuspended during stormy weather, or simply pass
through the canopy to the forest floor. Surface characteristics can influence foliar uptake,
chemical transformation, and resuspension. Landscape characteristics can affect wet
deposition via orographic effects and by the closer aerodynamic coupling to the
atmosphere of tall forest canopies as compared to the shorter shrub and herbaceous
canopies. The rainwater that passes directly through a canopy or is initially intercepted by
aboveground vegetative surfaces and subsequently drips from the canopy is measured as
throughfall. The fraction of the precipitation that drains from outlying leaves and
branches and is channeled to the stem of plants is classified as stemflow. Throughfall and
stemflow inputs constitute the majority of incident precipitation in forests and can
account for 70 to 90% of incoming precipitation in most cases, with the remainder lost to
interception within the canopy (Lcvia and Frost. 2003). Compared to wet deposition
measurements in the open, the magnitude of deposition from throughfall and stemflow
can either be smaller (e.g., from evaporation from canopy surfaces) or larger (e.g., from
resuspension of previously deposited material). The type of vegetation is important for
characterizing throughfall. For example, Templer et al. (2015a) found that the cycling of
N, particularly the rate of throughfall for NH44", is significantly different in conifer
compared to deciduous forest sites. Rainfall introduces new wet deposition and also
redistributes previously dry-deposited particles throughout the canopy. Intense rainfall
may contribute substantial total particulate inputs to the soil, but it also removes
bioavailable or injurious pollutants from foliar surfaces, while low-intensity events may
enhance foliar uptake through the hydrating of some previously dry-deposited particles
(U.S. EPA. 2004).
Chemistry within the canopy can also be important. Very fast measurements of NO2 flux
are confounded by the rapid interconversion of NO and NO2 with O3, and the biosphere
also interacts with NOx through hydrocarbon emissions and their subsequent reactions to
form multifunctional RONO2, including isoprene nitrates, which can account for a
substantial fraction of total N deposition. NO2 emissions also show UV dependence, and
both photo-induced and dark production of HNO2 from NO2 have been observed on leaf
surfaces, especially wet surfaces, although there is no consensus concerning chemical
mechanisms (U.S. EPA. 2008a). It was recently reported that NO2 deposition velocities
would have been overestimated by up to 80% if NO2 photolysis had not been considered
2-52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
(Brcuningcr et al.. 2012). Both biotic interactions and reaction chemistry further
complicate our understanding of N deposition and our ability to estimate deposition
velocity.
Some species, most prominently HNO3, can be characterized by unidirectional exchange,
whereas bidirectional exchange is more appropriate for most other species. Bidirectional
exchange is often described in terms of a compensation point, defined as the ambient
concentration above which a net uptake of the trace gas occurs and below which the trace
gas is released (Ganzcvcld et al.. 2002). A two-pathway process description can be used
to describe bidirectional exchange in a forest canopy [see e.g., (Fowler et al.. 2009;
Loubet et al.. 2001)1: (1) a stomatal pathway, which is bidirectional and modeled using a
stomatal compensation point and (2) a plant surface pathway, which denotes exchange
with water or waxes on the plant surface.
Understanding the exchange of reactive nitrogen species in the forest canopy has always
been challenging. One of the most comprehensive studies focusing on this question has
been the Biosphere Effects on AeRosols and Photochemistry Experiment (BEARPEX)
conducted in 2009 (Min et al.. 2014) that examined fluxes and transformations of NOx
within a forest canopy on the western slope of the Sierra Madre Range in 2009. The
study's results, shown schematically in Figure 2-10. indicate the existence of active
chemical interactions within the forest canopy in which NO emitted from soil or
transported from elsewhere is oxidized to NO2 and then to peroxy and alkyl nitrates and
HNO3. These pathways represent alternative mechanisms to plant uptake that have the
net effect of reducing the soil NO that escapes the forest canopy as NOx is converted to
peroxy nitrates and alkyl nitrates that can be transported to the atmosphere above the
canopy on very short time scales (~100s of seconds). The fraction of NO emitted by soil
that can be lost to the atmosphere above the canopy depends on the relative time scales
for transport through the canopy versus chemical transformation and foliar uptake.
Likewise, NO or NO2 transported form elsewhere can also be oxidized to organic nitrates
within the canopy. The organic nitrates formed can either be taken up within the canopy
or transported upward through the forest canopy to act as reservoirs of NOx that can
reform downwind.
2-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
HN03 = nitric acid; hv = solar photon; OH = hydroxyl radical; NO = nitric oxide; N02 = nitrogen dioxide; NOx = sum of NO and N02;
03 = ozone; RO = alkoxy radicals; R02 = organic peroxy radicals; R0N02 = alkyl nitrates; R02N02 = peroxy nitrates.
Bold arrows in blue (downward) and red (upward) represent the direction of the flux of each species across the canopy surface. Red
thin arrows within canopy indicate the nitrogen oxides removal processes within the canopy in addition to plant uptake.
Source: Min et al. (2014).
Figure 2-10 Schematic of the interactions involved in the exchange of
nitrogen oxides between the atmosphere and the forest canopy
as identified by Min et al. (2014).
The organic nitrates consist in large measure of isoprene- and monoterpene-derived
nitrates. Nguven et al. (2015) measured fluxes of organic nitrates on a tall tower in the
Talladega National Forest (AL) in June 2013 as part of the Southern Oxidant and Aerosol
Study (SOAS). They found that fluxes of organic nitrogen fonned by reactions of
nitrogen oxides (NO. NO2, NO3) with isoprene and monoterpene oxidation products
constituted -15% of the flux of oxidized N to the forest canopy, with most of the rest
from HNO3.
2.5.2.3. Bidirectional Exchange of NH3
NH3 can also be both emitted and deposited from plants and soils in a bidirectional
exchange. Farquhar et al. (1980) observed the existence of a compensation point for
ammonia due to gas exchange through the stomata of leaves. NH3 in the stomata results
from dissociation equilibria of NH4 produced physiologically in the leaves, followed by
equilibrium partitioning into air in the stomata (Sutton et al.. 1998). Further research
indicated that NH3 deposition rates to leaf surfaces were often faster than stomatal uptake
(Sutton et al.. 1993) and that NH3 can both react to form particulate NH4 and evaporate
from deposited PM within the canopy (Nemitz et al.. 2004; Brostetal.. 1988). Moisture
and plant type are strong influences, because deposition is more efficient on wet surfaces
(Sutton et al.. 1995). evaporation occurs under drying conditions (Fowler et al.. 2009).
June 2018
2-54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
and plant emissions are controlled by the physiological importance of NH44" in
intercellular fluids of plants (Sutton et al.. 1998). Copollutants also play a role because
NH3 deposition is enhanced by the presence of atmospheric acids like SO2 (Sutton et al..
1995: Mcleod etaL 19901
The complex pattern of NH3 sources and sinks with strong horizontal gradients of NH3
concentration (Fowler et al.. 2009) presents problems for simple bidirectional exchange
models (Sutton et al.. 1998). These problems have been addressed through the concept of
a canopy compensation point to include exchange with leaf surfaces (Flechard et al..
1999; Sutton et al.. 1998). as well as decomposing leaf litter and soil surfaces (Nemitz.
2000) in addition to stomatal exchange to describe bidirectional exchange (Burkhardt et
al.. 2008; Sutton etal.. 1998). Emission flux is particularly high from recently fertilized
soils (Fowler etal.. 2009; Sutton et al.. 1998) and after leaf-cutting events (Nemitz et al..
2009).
According to the 2008 ISA (U.S. EPA. 2008a). large areas of the U.S. are very near the
NH3 compensation point for most of the year, resulting in a highly dynamic air-surface
flux, which is prone to shifts in magnitude and direction. Bidirectional NH3 fluxes with
some periods of deposition and some periods of emission are typical for fertilized and
grazed agricultural ecosystems, while forests and other unfertilized ecosystems are
usually sinks for NH3 (Fowler et al.. 2009). Smaller emissions can also occur in semi
natural ecosystems (Fowler et al.. 2009).
Recently, regional scale modeling studies began to include canopy compensation points
and parameterize bidirectional exchange (Dennis et al.. 2010; Kruit et al.. 2010). and a
bidirectional exchange model for NH3 based on observations from North Carolina field
sites (Walker et al.. 2013) was developed for the CMAQ modeling system and an
agroecosystem model was included in CMAQ Version 5.0 to estimate NH3 emissions,
transport, and deposition from agricultural practices (Bash et al.. 2013). Including
bidirectional exchange in deposition modeling substantially improves agreement between
modeling results and ambient observations. A large bias of-19% has been observed in
annual wet deposition of NH44" when modeling results were compared with ambient
measurements without bidirectional exchange included in the model (Appel et al.. 2011).
NH4+ was underestimated throughout the year, but the largest underestimations were for
winter and spring in the Eastern U.S. (Appel et al.. 2011). The NH4+ wet deposition bias
was reduced by a factor 3, from -19 to -6%, by including bidirectional exchange in
CMAQ (Appel et al.. 2011).
Poor temporal and spatial representation of NH3 emissions in areas with fertilizer
application was also identified as a source of bias (Appel et al.. 2011). When CMAQ was
coupled with the U.S. Department of Agriculture's (USD A) Environmental Policy
2-55
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Integrated Climate (EPIC) agroecosystem model to improve characterization of fertilizer
emissions in annual simulations, NH3 dry deposition decreased by 45%, total NHx
deposition decreased by 15%, and total N deposition decreased by 5% compared to
modeling without bidirectional exchange (Bash et al.. 2013). In sensitivity tests of key
parameters in dry deposition modeling, the largest uncertainty was observed for the
change of unidirectional to bidirectional flux, but uncertainties of 5% or less in total
nitrogen deposition were reported (Dennis et al.. 2013). Although this is a small
difference nationwide, changes can locally be up to 50%, and only 66% of the 12 x 12
grid cells modeled showed changes of less than 10% (Dennis et al.. 2013).
Recent modeling studies have also improved insight into local areas and conditions under
which bidirectional flux most strongly affects deposition estimates. For example,
accounting for bidirectional flux resulted in a 17% increase in NH3 emissions from
agricultural operations (Massad et al.. 2010) compared with a 5% increase in
domain-wide NH3 emissions (Dennis et al.. 2013). Increases in NH3 emissions from
including bidirectional flux in semi natural ecosystems mostly occurred in areas of the
western U.S. with low emissions, where emissions were not included in existing
inventories. Seminatural ecoystems in the eastern U.S. isolated from agricultural
emissions exhibited changes of less than 1%. Seasonal differences were also observed,
with greater NH3 emissions observed in summer and winter, but emissions up to 45%
lower in fall and in spring when bidirectional exchange was included (Bash et al.. 2013).
2.5.2.4. Transference Ratios
Ratios of modeled or measured concentrations of SOx and NOy to their deposition, or
transference ratios (TRsox, TRnoy) were proposed by Scheffe etal. (2011) as a means to
link ambient air quality to deposition, and can be extended to NHx Transference ratios for
NOy, SOx, and NHx are given by:
•	TRnoy = (annual wet + dry deposition of NOyVannual average ambient
concentration of NOy
•	TRsox = (annual wet + dry deposition of SOx)/annual average ambient
concentration of SOx
•	TRnhx = (annual wet + dry deposition of NHx)/annual average ambient
concentration of NHx
These ratios are expressed in units of distance/time (as a velocity). In the 2011 Policy
Analysis for the NOx/SOx NAAQS review (U.S. EPA. 2011a). the transference ratios
were multiplied by measured ambient concentrations of NOy and SOx to estimate a flux.
June 2018
2-56
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
The transference ratio is an aggregate of multiple forms of nitrogen or sulfur. For
example, the transference ratio for NOy includes NO2 and HNO3, which have very
different deposition rates, chemical reaction rates, and atmospheric lifetimes. Close to
emission sources where fresh NOx emissions have had little time to react, the ratio of
HNO3 to NO2 is smaller than farther from emission sources. Accordingly, there is
considerable spatial and temporal variability (Sickles et al.. 2013) in transference ratios,
in part governed by the relative abundance of compounds with short and relatively longer
atmospheric lifetimes.
In practice, a regional modeling approach using a modeling system such as CMAQ or
CAMx is used in the calculation of transference ratios by simulating the relevant
transport processes discussed in Appendix 2.5.1. Sickles and Shadwick (2013) estimated
that TRsox and TRnoy could be given to within 25-35% of observed values using
observations of atmospheric concentration and deposition at some monitoring sites in the
eastern U.S. The study noted that accounting for year-to-year variability in precipitation
could lower the uncertainty. The transport processes described in Appendix 2.5.1 imply
that wet deposition should not necessarily be well correlated with surface concentrations
due to differences in the direction or spatial extent of transport in the boundary layer
compared to cloud levels. Dry deposition fluxes are more directly related to surface
concentrations. In a follow-on study, Sickles et al. (2013) calculated transference ratios
using measurements from CASTNET and NADP monitoring networks with CMAQ
model results and found the relative difference ranged from -37 to 64%. The authors
caution that this range should not be considered a definitive assessment of uncertainty
because relative differences do not reflect the extent to which the monitor is
representative of the grid-cell average modeled by CMAQ.
Koo et al. (2012) and Koo et al. (2015) raised the issue of model dependence on the
calculation of depositing species and transference ratios. Koo et al. (2012) compared
model results for concentrations of SO2, SO42 . HNO3, and NO3 and corresponding dry
deposition fluxes from CMAQ and CAMx to those measured at CASTNET sites. They
also compared model results for wet deposition to NADP/NTN measurements. On an
annualized basis, mean normalized errors (MNEs) in gas-phase concentrations ranged
from -25 to -100%. MNEs in dry deposition were much larger and ranged from -50 to
>300% and MNE in wet deposition ranged from -40 to -100% with no clear preference
for one model over another. MNE for NH4 in dry and wet deposition ranged from -35 to
70%. Koo et al. (2012) also found evidence for spatial variability in TRsox and TRnoy
across the U.S. and within selected ecosystems (roughly a few hundred km across).
Koo et al. (2015) compared simulations of transference ratios computed using CMAQ
and CAMx for two model years, 2005 and 2014, (see Figure 2-11) and found that TRsox
2-57
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
was much higher in CMAQ than in CAMx: however, differences were much smaller for
TRnoy- R2 values for TRnoy between the two models was 0.37 for 2005 and 0.33 for
2014. For TRsox, R2 was 0.073 for 2005 and 0.072 for 2014. Note that each point shown
in Figure 2-11 represents the average over 365 x 24 model entries and are dissimilar to
measurement artifacts. The outlying point (in the lower right) is from the Weminuche
Wilderness IMPROVE site (WEMI) in the Rockies and likely is the result of the
difficulty mesoscale models have in simulating precipitation and flow patterns in areas of
high relief and indicates further work is needed in this regard. Note that the disagreement
in TRsox by CMAQ is mostly due to an error in CMAQ emissions of SO2 that has since
been corrected. There is no consistent geographic pattern of agreement or disagreement
between the two simulations.
A more complete understanding of the causes of differences between model simulations
requires understanding the differences in how major chemical and physical processes
have been parameterized, thus underscoring the importance of accurately representing
emissions, transport, chemistry, and deposition. Both models used different modules to
represent these processes. In addition, these results might imply that the metrics used
need further scrutiny. Note again that the results shown in Koo et al. (2012) and Koo et
al. (2015) were obtained using older versions of CMAQ and CAMx and that CTMs are
continually undergoing improvement. In this regard, detailed comparisons with
observations and intercomparisons between the most current versions of these models
might help explain these findings.
Koo et al. (2015) also found very small differences between simulations for model years
2005 and 2014 from either CMAQ or CAMx for both NOy and SOx, indicating the ratios
are relatively invariant at least over an annual time scale. This result is not surprising,
because with the long averaging time, concentrations and deposition rates can better track
emissions changes.
2-58
DRAFT: Do Not Cite or Quote

-------
o 2005
+ 2014
o 2005
+ 2014
o CAMx
+ CMAQ
O CAMx
+ CMAQ
(a) TRj1[d 50x [CMAQ vs. CAMx)
{b)TRJtl0 SOx (2014vs. 2005)
(c) TR„io NOv (CMAQ vs. CAMx)
1.4
(dJTR^NOy [2014 vs. 2005)
1.4
1.5 2
CAMx
2005
CAMx
2005
CAMx = Comprehensive Air Quality Model with Extensions; CMAQ = Community Multiscale Air Quality; NOY = oxides of nitrogen;
SOx = sulfur oxides; TRatio = transference ratio.
Source: Adapted from Koo et al. (2015).
Figure 2-11 Scatterplots showing transference ratios for oxidized nitrogen
and sulfur oxides comparing the Community Multiscale Air
Quality model to the Comprehensive Air Quality Model with
Extensions in (a) and (c) and comparing 2005 to 2014 in (b) and
(d).
1	To summarize, recent studies have found that using transference ratios for estimating the
2	deposition flux from atmospheric measurements has lower uncertainty when applied at an
3	annual timescale. The transference ratio can vary spatially, and an estimate of uncertainty
4	and variability depends on the spatial scale of interest. Finally, previous studies have
5	highlighted some of the uncertainties when using models to calculate the transference
June 2018
2-59
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
ratio, but a comprehensive assessment of uncertainty is not available. When the
transference ratio is calculated using observations of atmospheric concentrations from
monitoring networks, the deposition flux for oxidized sulfur and oxidized nitrogen can be
estimated to within 25 to 35%.
2.5.3. Model Evaluation and Uncertainty
CMAQ model estimates were recently compared to monitoring network observations and
results were reported as normalized mean bias statistics (U.S. EPA. 2011a. 2009c). Total
nitrate concentrations were overestimated by CMAQ, with predictions averaged over
each of 4 years ranging from 22 to 26% higher than observed concentrations (U.S. EPA.
2011a). Model performance was described as good for total SOx, but the ability to
partition SOx into sulfate and SO2 was identified as an area that needed improvement
(U.S. EPA. 2011a). SO2 concentrations were overestimated, with CMAQ predictions
ranging from 39 to 47% higher than observed concentrations. Sulfate concentrations were
underestimated, with CMAQ predictions ranging from 9 to 17% lower than observed
concentrations (U.S. EPA. 2011a). A disadvantage of this type of comparison is that
modeled concentrations are outputted for a 12 * 12 km grid, while measured
concentrations are from a single point within that grid.
Wet deposition estimates from CTMs were extensively evaluated using wet deposition
data collected over the U.S. as part of the National Atmospheric Deposition Program's
National Trends Network. Simon et al. (2012) have summarized model evaluation studies
published before 2010, and Table 2-7 summarizes the studies since 2010.
While measurements comparing model outputs to observations provide one perspective
on the uncertainty in the fate and transport of atmospheric N and S, another approach to
quantifying uncertainty is to estimate the sensitivity of the model results with respect to
the uncertain range of parameters relevant to deposition calculations. A study by Dennis
et al. (2013) examined a range of uncertainties relevant to dry deposition using the
CMAQ model. This study found little change (<5%) in total deposition, despite changes
in dry deposition parameters because competing processes in the model tended to
rebalance and compensate. Changes in a single grid cell were as large as 20%.
There are also structural uncertainties that are difficult to assess in applying CTM models
to estimate deposition. The main structural uncertainties are associated with canopy
effects of NOx, bidirectional exchange of NH3, and transference ratios that relate average
concentration to deposition. These factors are discussed in Appendix 2.5.2.
June 2018
2-60
DRAFT: Do Not Cite or Quote

-------
Table 2-7 Reported comparisons of chemical transport models and
observations of nitrogen and sulfur wet deposition


Model
Metric
Accel et al. (2011)
7.9% normalized mean bias
CMAQ
Annual, wet deposition, sulfate,
continental U.S.
Accel et al. (2011)
-12.8% normalized mean bias
CMAQ
Annual, wet deposition,
ammonium, continental U.S.
Accel et al. (2011)
-12.8% normalized mean bias
CMAQ
Annual, wet deposition,
ammonium, continental U.S.
Accel et al. (2011)
-15% normalized mean bias
CMAQ
Annual, wet deposition, nitrate,
continental U.S.
Accel et al. (2011)
0.5 kg/ha median bias
CMAQ
Annual, wet deposition, nitrate,
continental U.S.
Zhanq et al.
(2012a)
6.5% mean normalized bias
GEOS-Chem
Annual, wet deposition, sulfate,
continental U.S.
Zhanq et al.
(2012a)
10% mean normalized bias
GEOS-Chem
Annual, wet deposition, nitrate,
continental U.S.
Zhanq et al.
(2012a)
7.4% mean normalized bias
GEOS-Chem
Annual, wet deposition,
ammonium, continental U.S.
Koo et al. (2012)
45 to 99% normalized mean error
CMAQ
Annual, wet deposition, sulfate,
continental U.S.
Koo et al. (2012)
38 to 99% normalized mean error
CMAQ
Annual, wet deposition, nitrate,
continental U.S.
Koo et al. (2012)
45 to 66% normalized mean error
CMAQ
Annual, wet deposition,
ammonium, continental U.S.
Williams et al.
(2017a)
0.34 kg ha-1 mean error
CMAQ
Annual, wet deposition, inorganic
nitrogen, Pacific Northwest
2.6. Geographic Distribution of Concentration and Deposition
1	Maps of national distributions of emissions, atmospheric concentrations, and deposition
2	fluxes of relevant species are presented in this section. The first two sections are limited
3	to deposition maps and are intended provide a broad overview of the extent and recent
4	trends for acid deposition (Appendix 2.6.1) and total nitrogen deposition, including
5	relative contributions of reduced and oxidized nitrogen (Appendix 2.6.2). The subsequent
6	three sections contain data on geographic distributions of emissions, ambient
June 2018
2-61
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
concentrations and deposition of oxidized nitrogen (Appendix 2.6.3). reduced nitrogen
(Appendix 2.6.4). and sulfur oxides (Appendix 2.6.5). including key species in each of
these classes.
Emission maps in Appendix 2.6.3. Appendix 2.6.4. and Appendix 2.6.5 are from an
earlier version of U.S. EPA's National Emissions Inventory than described in
Appendix 2.2. Ambient concentration maps in these sections are from a variety of
sources, depending on the availability of data.
Deposition maps in Appendix 2.6.1. Appendix 2.6.2. Appendix 2.6.3. Appendix 2.6.4.
and Appendix 2.6.5 are based on the approach of Schwede and Lear (2014a) which
combines measured and modeled values to produce spatially aggregated maps of wet,
dry, and total (wet plus dry) deposition of nitrogen and sulfur species across the U.S. Wet
deposition is based on concentrations measured in rainwater collected at NADP/NTN
monitoring sites combined with precipitation estimates interpolated by PRISM
(Parameter-elevation Regression Slopes Model) using inverse distance weighting (IDW).
In their approach to dry deposition, Schwede and Lear (2014a) measured values of
species concentrations in air at monitoring site locations and used bias-corrected
modeling results from CMAQ (currently at 12-km horizontal resolution) to fill in gaps
between sites and provide composition and deposition information for species not
measured (PANs, NO2, and HONO) in the routine monitoring networks. Distributions of
species that undergo dry deposition are derived mainly by fusion of data from the Clean
Air Status and Trends Network (CASTNET), the National Atmospheric Deposition
Program (NADP), Ammonia Monitoring Network (AMoN), and the Southeastern
Aerosol Research and Characterization (SEARCH) network. Parameterizations in CMAQ
are used to calculate deposition velocities for gases and particles (Pleim and Ran. 2011).
Note that CMAQ includes bidirectional exchange for NH3 (Bash et al.. 2012). but not for
other species such as NO2. Dry deposition fluxes are then combined with wet fluxes to
estimate total deposition. Details on interpolation, special treatment for particulate
species calculations, and procedures to correct for bias are described in detail by Schwede
and Lear (2014a).
Efforts have also been made to achieve greater consistency by relying more heavily on
WRF/CMAQ simulations for estimating wet deposition. Effects of biases in CMAQ wet
deposition are corrected by adjusting the modeled wet deposition by the ratio of observed
precipitation interpolated by PRISM to WRF precipitation. In this approach, it is assumed
that the ratio of observed to modeled precipitation is well correlated with the ratio of
observed to modeled wet deposition, but not (necessarily) that wet deposition scales
linearly with precipitation (Appel etal.. 2011). Likewise, estimates of dry deposition
could be obtained using CMAQ evaluated by comparison with monitoring results.
2-62
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
This approach resulted from efforts to improve estimates of atmospheric deposition by
advancing the science of measuring and modeling atmospheric wet, dry, and total
deposition of atmospheric species. Recognizing that the thin geographic coverage and the
lack of species measurements described in Appendix 2.4.6 are not ideal for estimating
dry deposition on a national scale, the National Acid Deposition Program established the
Total Deposition Science Committee (TDEP) with this mission in 2011, with an initial
goal of providing estimates of total S and N deposition across the U.S. for use in
estimating critical loads and other assessments, where loading results in the acidification
and eutrophication of ecosystems (NADP. 2016). Following this hybrid approach to
mapping total deposition that combines measured and modeled values, measured values
are given more weight at the monitoring locations, and modeled data are used to fill in
spatial gaps and provide information on chemical species that are not measured by
routine monitoring networks. This effort provides continuous spatial and temporal
coverage of total deposition estimates in the U.S., something previously unavailable
(NADP. 2016)1.
Limitations to the TDEP approach are: (1) interpolation leads to a minimization of
extreme values and a lower than actual variability, (2) data are limited to sites that meet
network completion criteria, (3) discontinuities in trends can occur for intermittent
monitoring data, (4) characterization of wet and dry organic nitrogen components is
uncertain and likely incomplete, (5) deposition in urban areas is not well represented
because the monitoring sites used are primarily in rural areas, and (6) occult deposition is
not well understood and might not be characterized accurately. An additional potential
drawback is that a mass balance is not maintained, although the mass balance error was
small in a similar effort combining measured wet deposition and bias corrected modeled
deposition (Schwedc and Lear. 2014a).
Differences in wet deposition of NH/, NOs . and SO42 and N + S expressed as H+
equivalents between the two, 3-year periods 1989-1991 and 2012-2014 across the U.S.
are shown in Figure 2-13. Figure 2-25. Figure 2-31. and Figure 2-38. These figures are
based on data obtained by the NADP/NTN. The maps were constructed by summing
gridded values and then taking the difference between the two, 3-year averages using data
from the NADP website. Although instructive, these results should be viewed with some
caution, as errors are incurred because development of a map spanning the CONUS
requires extensive interpolation between monitoring sites, which are often distant from
each other. The TDEP values shown on the maps are derived from NADP/NTN wet
deposition measurements at 4-km resolution coupled with CMAQ results for dry
deposition at 12 km. Weathers et al. (2006) found evidence for substantial spatial
variability in deposition with altitude, generally at smaller scales than used for TDEP. In
particular, they found evidence for factors of 4-6 variability in N and S deposition
2-63
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
throughout Acadia National Park (121 km2) and Great Smoky Mountains National Park
(2,074 km2); deposition rates averaged over the two parks were -70% higher than
inferred using only in situ point measurements. As might be expected, deposition
increased with elevation because of the increased importance of cloud deposition and dry
deposition (due to higher winds and hence increased turbulence). As noted by the NADP,
uncertainty within maps of wet deposition varies geographically and has not been
quantified. These maps are meant to provide a general indication of large-scale features
in the patterns and long-term changes in deposition.
Appendix 2.6.6 and Appendix 2.6.7 summarize concentration and deposition data from
other approaches. Appendix 2.6.6 describes distributions of dry deposition of NO2 and
SO2 from satellite-based measurements of tropospheric vertical column abundance and
model input derived by Nowlan et al. (2014). Appendix 2.6.7 describes recent estimates
of background concentrations, deposition fluxes, and sources and methods used to obtain
them.
Additional maps on the portion of the NADP website dedicated to the Total Deposition
(TDEP) program (http://nadp.sws.uiuc.edu/committees/tdep/tdepmaps/') are shown in
Appendix 2/7. They present a comprehensive overview of changes in various parameters
related to deposition from 2000-2013. Each map shows two 3-year periods, 2000-2002
and 2011-2113. Although uncertainty has not been fully characterized using the TDEP
approach, they are instructive because they give an indication of how various estimates
carried out with the same approach have changed over the past decade.
2.6.1. pH and H+ Equivalents
Long-term trends in rainwater pH over the CONUS between the two periods 1989 to
1991 and 2012 to 2014 are shown in Figure 2-12. Substantial improvement in the quality
of rainwater in terms of pH has occurred from the earlier to the later period. Figure 2-12
is a remarkable demonstration of the effectiveness of the Clean Air Act Amendments,
showing that the steep reductions of NOx and SO2 emissions described in Appendix 2.2
coincide with a sharp decline in pH.
June 2018
2-64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Source: NADP/U.S. EPA/CAMD.
Figure 2-12 (Left) pH of rainwater, 1989-1991; (Right) pH of rainwater,
2012-2014.
However, there are areas, especially in the eastern U.S. near the upper Ohio River Valley,
where the pH of rainwater is still much lower than the reference pH of -5.65 (for
equilibrium with CO2) that has often been used as a benchmark when characterizing the
excess acidity of rainwater. (Note that the reference CO2 concentration used to determine
this value is 316 ppm, atmospheric CO2 concentrations are now over 400 ppm, resulting
in a lowering of pH by -0.05 units.) As noted by Galloway et al. (1976). the major
contributors to free acidity in rainwater for pH <5.6 are the strong mineral acids HNO3
and H2SO4. However, weak acids (e.g,, organic acids) can contribute substantively to free
acidity at pH levels seen throughout much of the U.S. (see Appendix 2.3.5). For example,
concentrations of formic acid and acetic acid (pKa = 3.75, 4.76) measured in rainwater at
pH -5 by A very et al. (2006) are on the order of 10 uiVI. which is comparable to
concentrations of NO3 and SO42 measured in rainwater. Additionally, in areas like the
Northwest where the pH of stream water can be around or even larger than 5.6,
acidification of streams by CO2 might also need to be considered (Qu et al.. 2015).
The change in acid loading (H equivalents) due to wet deposition of NO;, . NHj . and
SO42 ions in precipitation expressed as H+ equivalents between the two, 3-year periods
1989-1991 and 2012-2014 across the U.S. based on data obtained by the NADP/NTN is
shown in Figure 2-13. Substantial decreases in acid loading are seen in the eastern U.S.,
with areas in the central and western U.S. showing smaller positive or negative changes
or essentially no change.
June 2018	2-65	DRAFT: Do Not Cite or Quote

-------
Change in
•#
Change in N and S Deposition (Eq. H+/ha) for 3-Yr. Avgs. Between 1989-1991 and 2012-2014
0	400	800 ^
Source: NADP NTN Annual Gradients
Nitrogen and Sulfur Wet Deposition from Nitrate, Ammonium,
and Sulfate in the U.S. between 1989 to 2014
H+ = hydrogen ion; eq = HH equivalent.
Figure 2-13 Difference in wet deposition of nitrate, ammonium, and sulfate
expressed as hydrogen ion equivalents (eq/ha/yr) over the
contiguous U.S. between 1989 to 1991 and 2012 to 2014.
1	This map was constructed by summing gridded values and then taking the difference
2	between the two, 3-year averages using data from the NADP website. Although
3	instructive, these results should be viewed with some caution, as errors are incurred
4	because development of a map spanning the CONUS requires extensive interpolation
5	between monitoring sites, which are often distant from each other. As noted by the
6	NADP, uncertainty within maps of wet deposition varies geographically and has not been
7	quantified. These maps are meant to provide a general indication of large-scale features
8	in the patterns and long-term changes in deposition. In viewing maps such as these one
9	should bear in mind that the rates and patterns of deposition are continually changing due
June 2018
2-66
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
to ongoing implementation of control measures and shifting patterns of population
growth, and industrial and agricultural activities.
2.6.2. Total Nitrogen
Figure 2-14 shows total deposition of N, which is given by the sum of all NOy and NHx
species considered by CMAQ. Fluxes are based on the method developed by Schwede
and Lear (2014a) as outlined above and are given in terms of kg N/ha/yr.
As can be seen from Figure 2-14. the highest deposition of nitrogen occurs in a broad
swath across the Midwest, and in more localized patches across the U.S. Total N
deposition can be put in context by comparing the deposition amounts in Figure 2-14 to
estimates of critical load, which is the N deposition amount below which no significant
harmful effects on sensitive elements of the environment occur. These are typically
below 10 kg N/ha/yr and can be as low as 2-3 kg N/ha/yr in both eastern and western
locations (Lee et al.. 2016; Ellis et al.. 2013). amounts which are firmly below the
estimated deposition amounts in Figure 2-14 over wide areas of the U.S. Ellis et al.
(2013) observed that critical loads were exceeded in 24 of 45 parks, and Lee et al. (2016)
observed that critical loads were exceeded at more locations in the western U.S., but by
larger amounts in the eastern U.S.
Inspection of Figure 2-15a and b shows that many of these areas are dominated either by
deposition of NHx emitted mainly by agriculture (e.g., California's Central Valley, Upper
Midwest), or NOy resulting from oxidation of nitrogen oxides emitted mainly by
combustion sources (e.g., the Northeast, Southwest). As might be expected when
considering all forms of NOy, total deposition tends to be much higher near urban and
suburban areas. Note also that deposition of Nr, based on the hybrid
measurements/modeling approach, is dominated by reduced forms across the CONUS as
a whole.
June 2018
2-67
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
Total deposition of nitrogen 11L3
Soiure: CASTNFT/CMAQ/NTN'/AMON/SH ARCH	TTSHPA IOWI4
N = nitrogen; Nr = reactive nitrogen.
Source: CASTNET/CMAQ/NTN/AM O N/S E ARC H.
Figure 2-14 Total (wet + dry) deposition of nitrogen (kg N/ha/yr) over the
contiguous U.S. 2011-2013.
Dry deposition of gas-phase N (as HNO3 and NH3) exceeds dry deposition of particulate
forms (pNO , . N H4 ) over most of the CONUS according to the hybrid method. Overall,
deposition of N is mainly as reduced forms, with a maximum over the north-central U.S.
The Central Valley of California, northern Utah, and eastern North Carolina are among
other areas of high deposition of reduced N. In general, dry deposition of N, in either
oxidized or reduced form, exceeds wet deposition across the CONUS. However, based on
discussions in earlier sections, uncertainties for dry deposition are likely much larger than
for wet deposition.
As mentioned earlier, several species potentially important for deposition are not
measured in CASTNET. Figure 2-16 shows dry deposition for oxidized nitrogen species
(e.g., PAN, other organic nitrates, HONO) calculated by CMAQ.
June 2018
2-68
DRAFT: Do Not Cite or Quote

-------
Pet of total N as reduced N 1113
Source: CASTNET/CMAQ/NTO/AMON/SEARCH	USEPA 10/16/1-4
reN = reduced nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH
Pet of total N as oxidized N 1113
Source: CASTNET/CMAQ/NTN/AMON/SEARCH	USEPA 10/16/14
oxN = oxidized nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-15 (A) Percentage of total nitrogen deposition as reduced inorganic
nitrogen over the contiguous U.S. 2011-2013. (B) Percentage of
total nitrogen deposition as oxidized nitrogen over the
contiguous U.S. 2011-2013.
June 2018
2-69
DRAFT: Do Not Cite or Quote

-------
Pet of total N as unmeasured species 1113
USEPA10/16/1-1
Source; CASTNfeT/CMAQ/NTN/AMOM.'SEARCH
Other N
(Pet ot Total)
N = nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-16 Three-year average percentage of total nitrogen deposition by
species (i.e., those species that are not measured in the
networks) simulated by the Community Multiscale Air Quality
modeling system.
1	As seen in Figure 2-16. deposition of these species can contribute substantially to N
2	deposition, especially near strong sources, in particular large urban areas. Turnipseed et
3	al. (2006) also indicated that not accounting for these species can result in significant
4	underestimates of N deposition.
5	Figure 2-17 shows long-term trends in total wet deposition of N (MEET + NOa ) binned in
6	increments of 2 kg N/ha/yr for comparison to critical loads estimates. Although it is
7	apparent that N wet deposition has decreased overall across the U.S., there are areas
8	showing increases. Also shown are NTN sites active during either period.
June 2018
2-70
DRAFT: Do Not Cite or Quote

-------
Sum NO3' and NH4+ Wet Deposition by 3-Yr. Averages
2012 - 2014
1989 - 1991
t
• •

'3K W >./-V
* , • ••*!•• .

Kg N I ha /yr	• •
~ <2
J > 2 to 4
J > 4 to 6
| > 6 to 8
| > 8 to 10
¦¦ > 10
• Active NTN Sites '89-'91 or ,12-,14
0	500	1.000 ^
Source: NADP NTN Annual Gradients
NTN = National Trends Network: N = nitrogen; NH4H' = ammonium; N03" = nitrate.
Figure 2-17 Wet deposition of ammonium + nitrate (kg N/ha/yr) over the
contiguous U.S. in two, 3-year periods, 2012 to 2014 and 1989 to
1991. Also shown are active National Trends Network sites in
either period.
June 2018
2-71
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Studies at individual sites (e.g., on the coast of North Carolina) have shown that about
30% of wet deposition of N consisted of organic N, 20-30% of which was then available
to primary producers on timescales of hours to days (Pcicrls and Paerl. 1997). In addition,
Benedict et al. (2013) found that wet deposition of organic nitrogen contributed 18% of
total quantified reactive nitrogen deposition and 25% of wet nitrogen deposition at Rocky
Mountain National Park between November 2008 and November 2009. Trends in total
(wet + dry) deposition of total (NOy + NHx) nitrogen between 2000 and 2013 are
described with maps in Appendix 2/7.
2.6.3. Oxidized Nitrogen
County level NOx emissions are shown in Figure 2-18. The emission map is currently
being updated and Figure 2-18 shows estimated emissions from 2001. At present, areas
of high emissions that stretch over numerous counties are in the Northeast, Southern
Great Lakes region, and Florida. The distribution of ambient NOy concentrations is
shown in Figure 2-19. However, because NOy is only measured at a small number of
sites, this map is based solely on CMAQ model output. The distribution of NO2 (shown
in Figure 2-20) is derived from satellite data (OMI) and output from the GEOS-Chem
model using the method outlined in Appendix 2.6.3. Distributions of HNO3, pNOs .
pNH4+, SO2, and pSOr are based on data from the CASTNET. Note, however, that
because of artifacts relating to measurement of HNO3 and pNOs (Appendix 2.4.5). the
measurement of total nitrate (TN = HNO3 + pNOs ) is judged to be more reliable than
measurements of its components.
June 2018
2-72
DRAFT: Do Not Cite or Quote

-------
2001 County Emissions (1000 Tons per Year) of Nitrogen Oxides
0	>D-0.S4	0.54-1.2	1.2-2.5
I 2.5-6 fli 6-16	¦ 16+
Source: U.S. EPA (2008a).
Figure 2-18 Distribution of annual NOx emissions in 2001.
June 2018
2-73
DRAFT: Do Not Cite or Quote

-------
NOY
~ 33-86
N0y = oxidized nitrogen species.
Source: U.S. EPA/OAQPS.
Figure 2-19 Distribution of annual average total oxidized nitrogen species
concentrations for 2011 simulated by Community Multiscale Air
Quality modeling system.
June 2018
2-74
DRAFT: Do Not Cite or Quote

-------
OMI-derived surface N02 (ppb)
I
7.00
6.22
5.44
4.67
3.89
3.11
2.34
1.56
0.78
0.01

JJA
DJF = December, January, February; JJA = June, July, August; N02 = nitrogen dioxide; OMI = Ozone Monitoring Instrument.
Note: Images shown were constructed by Dr. Lok Lamsal of Universities Space Research Association from data obtained by the
OMI on the Aura satellite (http://aura.gsfc.nasa.gov/scinst/omi.html) using the algorithm described in Bucsela et al. (2013). Output
from the GEOS-Chem, global-scale, three-dimensional, chemistry-transport model to derive surface concentration fields from the
satellite data as described in Lamsal et al. (2008) and Lamsal et al. (2010).
Top panel (winter; December, January, February). Lower panel (summer; June, July, August).
Figure 2-20 Seasonal average surface nitrogen dioxide mixing ratios in parts
per billion for winter (upper panel) and summer (lower panel)
derived by the Ozone Monitoring Instrument/GEOS-Chem model
for 2009-2011. The Ozone Monitoring Instrument has an overpass
at approximately 1:30 p.m. local standard time.
June 2018
2-75
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
As seen in the figure, highest values are found in and around urban areas. In general,
much broader areas of high concentrations (>~5 ppb) are found in the eastern U.S., with
many areas in the western U.S. subjected to concentrations <1 ppb, which implies that
concentrations of components that might pose a hazard are also lower than 1 ppb because
NOy refers to the sum of oxidized N species.
Figure 2-20 through Figure 2-22 describe geographic concentration patterns for three
major oxidized nitrogen species, NOy, HNO3, and pNOs . Figure 2-20 shows seasonal
average NO2 concentrations derived using the hybrid (OMI-satellite/GEOS-Chem-model)
approach described in Appendix 2.4.2.2. Large variability in NO2 concentrations is
apparent in Figure 2-20. As expected, the highest NO2 concentrations are seen in large
urban regions, such as in the Northeast Corridor, and lowest values are found in sparsely
populated regions located mainly in the West. Minimum hourly values can be less than
-10 ppt, leading to a large range between maximum and minimum concentrations.
Although overall patterns of spatial variability are consistent with the current
understanding of the behavior of NO2, not much confidence should be placed on values
<~100 ppt due to limitations in the satellite retrievals. Surface NO2 concentrations tend to
be higher in January than in July, largely reflecting lower planetary boundary layer
heights in winter. Such seasonal variability is also evident on a local scale, as measured
by surface monitors. For example, in Atlanta, GA, NOx measurements also exhibited
higher concentrations in winter and lower concentrations in summer, when NOx is more
rapidly removed by photochemical reactions. For example, see U.S. EPA (2008b).
Figure 2-21 shows ambient concentrations of HNO3. Elevated concentrations of HNO3
are notable in southern California, the Midwest, the south-central U.S., and the
Mid-Atlantic states. Conversion of NO2 to HNO3 takes place over a timescale of 1 to
several hours, during which time appreciable transport can occur.
Figure 2-22 shows 3-year average concentrations of particulate nitrate (pNOs ) across the
CONUS. Average pNOs concentrations were highest in the Upper Midwest with a
notable maximum at the junction of Iowa, Wisconsin, Missouri, and Illinois. The high
values in the Upper Midwest are expected to be found during winter for reasons noted in
Appendix 2.3.3. Elevated levels were also observed in central California in the San
Joaquin Valley, central Pennsylvania, central Florida, and through much of the Midwest.
Based on IMPROVE and CSN monitoring network data, ammonium nitrate
concentrations are highest in California and the Midwest (Hand et al.. 2012c). It is worth
noting that at several monitoring sites in the central and northern Great Plains, NO3 and
S042 are increasing at a rate of over 5% per year (Hand et al.. 2012a). Hand et al.
(2012a) suggested that this increase might be related to oil and gas exploration and
2-76
DRAFT: Do Not Cite or Quote

-------
1
2
production in the region, transport from oil and gas fields in Alberta, and also to
expansion of EGUs to meet the demands of population growth.
HNO3
(fig/m3)
-0.0
-0.5
-1°
-1.5
-2.0
r
j-3.0
L3.5
—I->4.0
Source: CASTNET
HNO3 = nitric acid.
Concentrations of nitric acid (|jg/m3) can be converted to mixing ratios (parts per billion) to rough approximation at normal
temperature and pressure by multiplying by 0.38.
Source: CASTNET/U.S. EPA-CAMD 10/09/14.
Figure 2-21 Three-year average (2011-2013) surface concentrations of nitric
acid based on monitoring data obtained at Clear Air Status and
Trends Network sites (black dots).
USEPA/CAMD 10/09/14
,<:ta ta fa	1113 /kncJ _c-i 113
June 2018
2-77
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
N03 = nitrate.
Source: CASTNET/U.S. EPA-CAMD 10/09/14.
Figure 2-22 Three-year average (2011-2013) surface concentrations of
particulate nitrate based on monitoring data obtained at Clear Air
Status and Trends Network sites (black dots).
Source: CASTNET
USEPA/CAMD 10/09/14
.yateijr^czc.tiie^mg/l 113/noJ_e-i 113
Figure 2-23 shows the depositional flux of NO? over the CONUS 2011-2013. In
constructing Figure 2-23 and maps that include pNO, deposition estimates, the
assumption was made that 80% of pN(V is in the fine mode and 20% is in the coarse
mode. Note also, the monitors in CASTNET do not use size-selected inlets. As shown in
Figure 2-9 (Appendix 2.5.2.1). the deposition velocity of particles increases dramatically
with particle size due to gravitational settling, resulting in higher deposition rates than
calculated if pNO-; were found mainly in the fine mode. However, to the extent there is
displacement of FINOs by acid S species in fine particles as occurs in the eastern U.S. due
to much higher emissions of SO2 than in the western U.S. [e.g., (Wolff. 1984); especially
in coastal areas where displacement of Cr in marine aerosol occurs], higher levels of
pNO3 would be found in coarse mode particles (see Appendix 2.3.3. During the Tampa
Bay Study, measured and modeled (using CMAQ) concentrations of HNO3 and pNO? in
PM1Q-2..5 were much higher than pNO, in PM2 5. Wolff (1984) found that most pNO, (as
June 2018
2-78
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
NH4NO3) is found in the fine mode in Denver but in the coarse mode (associated with
Ca: and Mg2 ) in measurements made in Detroit and rural South Dakota, Louisiana, and
Virginia. Blanchard et al. (2013) found a range of 32 to 63% for the fraction of pNO; in
PM io ;,5 versus PM2 5 particles in the Southeast. Lee et al. (2008) found that most pNOV
was in the coarse mode at Grand Canyon and Great Smoky Mountains, corroborating
earlier findings at Yosemite and Big Bend national parks. They also found that both
coarse and fine mode pNOa were important at Brigantine National Wildlife Refuge, NJ
and San Gorgonio Wilderness Area, CA. Lefer and Talbot (2001) also found that NO,
sampled at Harvard Forest, MA between March and October was mainly found in the
coarse mode with a mass median diameter of 4.8 ±1.5 pm. These results indicate
considerable regional variability in the ratio of pN(K in the fine and coarse modes and
consequently additional uncertainty in estimates of pNO ; deposition.
Souire: C 'ASTNHnt:MAQ/NrrN/AMON/SKAK.( !H
Tota] deposition of oxidized N 1113
USEPA 10/16/14
Total oxN
(kg-N/ha)
oxN = oxidized nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-23 Total oxidized nitrogen deposition over the contiguous U.S.
2011-2013.
Figure 2-24 and Figure 2-25 show that most locations in the U.S. show decreases in wet
deposition ofNO-; . which are associated with NOx emissions control measures since the
passage of the 1990 Clean Air Act amendments. Flowever, some areas, located mainly in
June 2018
2-79
DRAFT: Do Not Cite or Quote

-------
1	the West show increases. As noted earlier, Hand et al. (2012a) suggested that this
2	increase might be related to oil and gas exploration and production in the region,
3	transport from oil and gas fields in Alberta, and also to expansion of motor vehicles and
4	EGUs to meet the demands of population growth. The large area with the strongest
5	increases in the north-central U.S. corresponds to oil and gas operations in the Bakken
6	Shale. Trends in total (wet + dry) deposition of oxidized nitrogen between 2000 and 2013
7	are described with maps in Appendix 2.7.
Source: NADP/U.S. EPA/CAMD.
Figure 2-24 (Left) nitrate wet deposition, 1989-1991; (Right) nitrate wet
deposition, 2012-2014.
June 2018
2-80
DRAFT: Do Not Cite or Quote

-------
Change in N03~ Wet Deposition in the U.S.
between (1989-1991) and (2012-2014)
Change in N03 (Kg N/ha/yr) for 3-Yr. Avgs. Between 1989-1991 and 2012-2014
** J* JV	J*
V	
-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
Figure 2-27 and Figure 2-28 show maps for the reduced inorganic nitrogen species, NH3
and pNH4+. The distribution of NH3 was obtained from the Ammonia Monitoring
Network (AMoN). Highest concentrations of NH3 were measured in Salt Lake City, UT
at 15 |ig/nr\ All other annual average concentrations for 2012 were lower than 5 |ig/nr\
The Salt Lake City site is located near feed lots, perhaps explaining in large part why
levels were much higher there than at other sites. In general, areas of highest NH3
concentration correspond well with areas of highest NH3 emissions, as shown in the 2008
ISA for Oxides of Nitrogen (U.S. EPA. 2008^). Note that confidence in the magnitude
and inter-monitor precision of NH3 measurements has increased since the 2008 ISA (U.S.
EPA. 2008a'). also see Appendix 2.4.3.1. However, sparseness of the monitoring network
still presents uncertainty in describing the nationwide distribution of NH3 concentrations.
Particulate NH4 concentrations were obtained from CASTNET, and were highest in
Illinois-Indiana-western Ohio, along with high values in central Pennsylvania and central
California. These locations correspond generally to highest concentrations of pNO;, and
moderate-to-high concentration locations for NH3. In addition, some of the areas with
high NH3 concentrations, such as southern Wisconsin or Salt Lake City, do not appear to
have elevated pNH/.
Figure 2-29 shows the depositional flux of NHx over the CONUS 2011-2013.
2-82
DRAFT: Do Not Cite or Quote

-------
2001 County Emissions (Tons per Year) of Ammonia
690-1300
>0—120
1300-2400
120-320
2400+
320-690
Source: U.S. EPA (2008a).
Figure 2-26 Distribution of annual ammonia (NHs) emissions by county in
2001.
June 2018
2-83
DRAFT: Do Not Cite or Quote

-------
Ambient Ammonia Monitoring Network (AMoN)
nh3
Concentration
(pg/m3)
O -= 0.40
O 0.40 - 0.80
O 0.80-1.20
O 1.20-1.60
O 160-2.00
O 2.00 - 2.40
O > 2.40
2012
Average Ambient
Ammonia Concentration
AMoN = Ambient Ammonia Monitoring Network; NH3 = ammonia.
Source: CASTNET/U.S. EPA-CAMD 10/09/14.
Figure 2-27 Average (2012) surface concentration of ammonia obtained by the
Ambient Ammonia Monitoring Network at select Clear Air Status
and Trends Network sites. Concentrations of ammonia (pg/m3)
can be converted to mixing ratios (parts per billion) to rough
approximation at normal temperature and pressure by multiplying
by 1.4.
June 2018
2-84
DRAFT: Do Not Cite or Quote

-------
Source: CASTNET
USEPA/CAMD 10/09/14
AlaB Ajri'caitiiet^mg/1113/nK4_c-ll 13
NH4+ = ammonium.
Source: CASTNET/U.S. EPA-CAMD 10/09/14.
Figure 2-28 Three-year average (2011-2013) surface concentrations of
particulate ammonium (pg/m3) based on monitoring data obtained
at Clear Air Status and Trends Network sites (black dots).
June 2018
2-85
DRAFT: Do Not Cite or Quote

-------
Total deposition of reduced N 1113
USEPA 10/16/14
Source: CASTNFT/CMAQ/NTN7AMON/SF.AR CH
Total reM
(kg-Ntoa)
reN = reduced nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-29 Total reduced inorganic nitrogen deposition over the contiguous
U.S. 2011-2013.
1	Figure 2-30 and Figure 2-31 shows large increases in wet deposition of NH4+ centered in
2	eastern North Carolina and throughout the north-central U.S. and decreases in the Gulf
3	States. The situation is more nuanced than shown in that some sites show small increases
4	and others small decreases. In general, large-scale increases in wet deposition of NFI/,
5	rather than decreases, are seen across the U.S. in agreement with the analysis of Ii et al.
6	(2016d). Trends in total (wet + dry) deposition of reduced nitrogen between 2000 and
7	2013 are described with maps in Appendix 2.7.
June 2018
2-86
DRAFT: Do Not Cite or Quote

-------
Source: NADP/U.S. EPA/CAMD.
Figure 2-30 (Left) ammonium wet deposition, 1989-1991; (Right) ammonium
wet deposition, 2012-2014.
June 2018
2-87
DRAFT: Do Not Cite or Quote

-------
Change in NH4+ Wet Deposition in the U.S.
between (1989-1991) and (2012-2014)
¦j*
,,./w
Change in NH4* (Kg N/ha/yr) for 3-Yr. Avgs. Between 1989-1991 and 2012-2014
mit
jv	^
V	N' V V
aO	tO	(O	tO	1
*9 *9 . hP  * K
Sourcs NADP NTN Annua! Grmfitf/vts
N = nitrogen; NH4+ = ammonium.
Figure 2-31 Difference in wet deposition of ammonium (kg N/ha/yr) over the
contiguous U.S. between 1989 to 1991 and 2012 to 2014.
2.6.5.	Sulfur Oxides
1	Figure 2-32 shows the west-to-east increasing gradient in SO2 emissions, with greater
2	emissions in most counties east of the Mississippi than in most counties in the West. The
3	emissions map is being updated, but geographic patterns have changed little. Widespread
4	areas of high emissions that encompass numerous counties are in the Northeast, the Ohio
5	Valley, and parts of the South and West.
June 2018
2-88
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
2001 County Emissions (1000 Tons per Year) of Sulfur Dioxide
0	>0-0.05 	 0.05-0.11	0.11-0.31
0.31-1.5 H 1.5-12	¦ 12+
Source: U.S. EPA (2008a)
Figure 2-32 Distribution of annual sulfur dioxide (SO2) emissions by county
from the 2014 National Emissions Inventory.
Figure 2-33 and Figure 2-34 show the distribution of atmospheric concentrations of
gas-phase SO2 and particulate phase SO42 . Elevated SO2 concentrations persist along the
Ohio River Valley and western Virginia, but concentrations have decreased substantially
over the last decade throughout the eastern U.S. Comparison between the national SO2
distributions (Figure 2-33) for 2011-2013 and the ones for 1989-1991 and 2003-2005
presented in the 2008 ISA (U.S. EPA. 2008a) demonstrated continual decreases in SO2
concentrations across the nation.
Both concentrations and seasonal variability of sulfate are substantially higher in the
eastern U.S. than in the West (Hand et al.. 2012c). Based on air pollution monitoring
network data (IMPROVE and CSN), sulfate concentrations on a national scale are
steadily decreasing across the U.S. Between 1992 and 2010, annual mean sulfate
concentrations at rural sites decreased fairly consistently at a rate of-2.7% per year. This
decline has become even steeper more recently, with annual mean concentrations
decreasing by an average of-4.6% per year from 2002 to 2010. The decrease appears to
be due to decreasing SO2 emissions from power plants (Hand et al.. 2012b). While the
nationwide trend is for a reduction in sulfate concentrations, there are seasonal and
June 2018
2-89
DRAFT: Do Not Cite or Quote

-------
regional increasing trends, specifically in the central and northern Great Plains in winter,
and in the western U.S. in spring (Hand et al.. 2012a). Both sulfate and nitrate are
increasing at a rate of over 5% per year at several monitoring sites in the central and
northern Great Plains (Hand et al.. 2012c).
Source: CASTNET	USEPA/CAMD 10/09/14
tdabafaizJca'Stnctfjmg/l 11 J/so2_o4113
S02 =sulfur dioxide.
Source: CASTNET/U.S. EPA-CAMD 10/09/14.
Figure 2-33 Three-year average (2011-2013) surface concentrations of sulfur
dioxide obtained by fusion of monitoring data obtained at Clear
Air Status and Trends Network sites (black dots) and Community
Multiscale Air Quality model system results. Concentrations
(pg/m3) can be converted to mixing ratios (parts per billion) at
normal temperature and pressure) to rough approximation by
multiplying by 0.37.
June 2018
2-90
DRAFT: Do Not Cite or Quote

-------
S042 = sulfate.
Source: CASTNET/U.S. EPA-CAMD 10/09/14.
Figure 2-34 Three-year average (2011-2013) surface concentrations of
particulate sulfate based on monitoring data obtained at Clean Air
Status and Trends Network sites (black dots).
Source: CASTNET
USEPA/CAMD 10/09/14
/data/arc'ai stnet/jaig/1 113/io4_c4113
1	Figure 2-35 shows wet plus dry deposition of SOx (SO2 + SO42 ) over the CONUS.
2	Maximum deposition occurs over the Ohio River Valley (southeastern Ohio, West
3	Virginia, and western Pennsylvania). Figure 2-36 shows a good deal of spatial variability
4	in the percentage of dry deposition across the CONUS. In the area of highest total
5	deposition in the Mid-Atlantic states, dry deposition of SO2 is dominant and dry
6	deposition of pS042 is very minor component of dry deposition of SOx. Wet deposition
7	dominates in the Pacific Northwest, northern New England and in general in the central
8	U.S. Note that deposition of organic sulfur species (e.g., methane sulfonic acid) or SO2
9	and/or SO42 produced by the oxidation of organic S species is not included.
June 2018
2-91
DRAFT: Do Not Cite or Quote

-------
Siiiinrn: CASTN I TlX-MAQ.'NTN.'AMOK/SliAU11 [
Towl 
-------
Pel of tola] S us thy deposition 1113
USHfA iCvl
-------
WetSCU2*
(tcg/ha)
Wet S042"
(kg/ha)
f
-12
-16
20
24
28
>32
			 		USETAfCAMDOMO'lfc
Source: JtADP/NTPi &PRKM	USEPA/CAMD03(02/12
Figure 2-37 (Left) sulfate wet deposition, 1989-1991; (Right) sulfate wet
deposition, 2012-2014.
Change in SO/' Wet Deposition in the U.S.
Change in SO*2 (Kg S/ha/yr) for 3-Yr. Avgs. Between 1989-1991 and 2012-2014
S = sulfur; S042 = sulfate.
Figure 2-38 Difference in wet deposition of sulfate (kg S/ha/yr) over the
contiguous U.S. between 1989 to 1991 and 2012 to 2014. The
range of positive values is much smaller than for negative values.
June 2018
2-94
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
The increases in wet deposition of SO42 in the north-central U.S. correspond to those for
N03 and are in the immediate vicinity of the Bakken Shale. There is a high degree of
interannual variability in deposition in some areas, especially those showing increases
(e.g., Logan, UT/Idaho), making source attribution difficult. Trends in total (wet + dry)
deposition of sulfur between 2000 and 2013 are described with maps in Appendix 2/7.
2.6.6. Particulate Matter (PM)
Figure 2-39 shows the 3 year mean of the 24-hour PM2 5 concentrations for PM2 5
network monitoring sites across the U.S. from 2013-2015. Emissions are not shown
because the majority of PM2 5 mass is often produced by atmospheric reactions (see
Appendix 23). Some of the highest PM2 5 concentrations are in the San Joaquin Valley
and the Los Angeles-South Coast Air Basin of California. However, in general 3-year
average 24-hour PM2 5 concentrations are higher in the Eastern U.S. than in the Western
U.S. An area of the highest concentrations in the Eastern U.S. can be seen in the Ohio
Valley. From Appendix 2.3.6 is the same area where SO42 and NO3 account for the
greatest fraction of PM2 5, indicating that at least in the Eastern U.S., the highest PM2 5
concentrations also correspond to PM with greatest fraction mass account for by SO42
and NC>3~.
June 2018
2-95
DRAFT: Do Not Cite or Quote

-------
Figure 2-39 3-year average concentrations of particulate matter smaller than
2.5 pm diameter (PM2.5) 2013-2015.
1	A further indication that the fraction of mass contributed by SO42 and NO, may increase
2	with increasing PM2.5 concentration is demonstrated by Figure 2-40. Figure 2-40 shows
3	that a decrease in national average PM25 between 2000 and 2015 is paralleled by a
4	similar decrease in SO:2 . Since SO42 has been the most abundant component of I'M
5	the steep decline in SO2 emissions (Appendix 2.2) has led to a sharper decrease in S042
6	concentrations than 111 concentrations of other PM2.5 components. As indicated by
7	Figure 2-39. in those areas where SO42 concentrations remain high, the S042 fraction of
8	PM2.5mass is also still high in the Eastern U.S.
June 2018
2-96
DRAFT: Do Not Cite or Quote

-------
Notes: back = mean, gray = 90th percentile.
Source: Chan et al. (2018).
Figure 2-40 National monthly average concentrations of particulate matter
smaller than 2.5 pm diameter (PM2.5; top) and sulfate in PM2.5
(bottom) from 2000-2015 (concentrations in pg/m3).
1	For completeness, Figure 2-41 shows PMa-t.5 concentrations. The highest concentrations
2	are observed in Southwest and Great Plains. These are the same areas where crustal
3	material accounts for the greatest fraction of PM2.5 as described in Appendix 2.3.6. and
4	PMie-2 5 is also largely composed of crustal material, which has little impact on acid or
5	nutrient deposition.
June 2018
2-97
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Source: EPA 2016 analysis of Air Quality System network data 2013-2015.
Figure 2-41 98th percentile concentrations for PM10-2.5 between 2013-2015.
2.6.7. Distributions of Dry Deposition of Nitrogen Dioxide and Sulfur Dioxide
Derived Using Satellite-Based Measurements and Chemistry Transport
Models
Figure 2-42 shows the annual average dry deposition velocities and fluxes of NO2 and
SO2 for 2005 to 2007 and their estimated uncertainties derived by Nowlan et al. (2014)
using data derived from the Ozone Monitoring Instrument (OMI) on board the Aura
satellite and model parameters from the GEOS-Chem three-dimensional,
chemistry-transport model.
As shown in Figure 2-42. higher fluxes for both NO2 and SO2 occur in the East than in
the West. In particular, there is a band of high dry deposition for NO2 and SO2 along the
Ohio River. High depositional fluxes for NO2 are also seen along the Northeast Corridor
and in scattered locations throughout the East. In addition, there is a noticeable plume of
SO2 over the western Atlantic Ocean. Average, relative uncertainty in the flux estimates
for both NO2 and SO2 are -30% over land and are not much higher over the Atlantic
Ocean south of Massachusetts and Nova Scotia. Increased SO2 deposition, especially
near shore, is expected based on the likelihood of off-shore transport of SO2 and NO2
along with other pollutants by synoptic weather systems. Note that bidirectional exchange
for NO2 (and a number of other gases) has not been implemented yet in GEOS-Chem or
June 2018
2-98
DRAFT: Do Not Cite or Quote

-------
1	in CMAQ. Note also that in this study, the algorithms used to derive NO2 and SO2
2	columns are older than more recent ones with lower detection limits. These results,
3	however, do illustrate the potential of the hybrid, satellite/model approach for mapping
4	deposition at the continental scale.
June 2018
2-99
DRAFT: Do Not Cite or Quote

-------
N02 Dry Deposition Flux [kg N ha"' yr'1]
S02 Dry Deposition Flux [kg S ha'1 yr'1]
0	0.1	0,2	>0,3	0 1 2 3 4 >5
N02 Dry Deposition Uncertainty [kg N ha"1 yr'1]	S02 Dry Deposition Flux Uncertainty [kg S ha'1 yr'1)
N = nitrogen; N02 = nitrogen dioxide; S = sulfur; S02 = sulfur dioxide.
Source: Nowian et al. (2014).
Figure 2-42 Top panel: modeled deposition velocities for nitrogen dioxide and
sulfur dioxide for 2005 to 2007; middle panel Satellite-model
estimates of annual mean dry deposition fluxes of nitrogen
dioxide and sulfur dioxide; bottom panel: uncertainties in
estimates.
June 2018
2-100
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
2.6.8.
Background Concentrations and Deposition
Background refers here to those concentrations or fluxes that do not result from U.S.
anthropogenic emissions. Background sources of N include natural sources like lightning,
wildfires, and emission from soils. Background sources of SOx include natural sources,
such as volcanos or oxidation of reduced sulfur species (H2S, | CH,12S) emitted in
anaerobic environments, and anthropogenic sources from outside the U.S. Background
levels so defined facilitate separation of pollution levels that can be controlled by U.S.
regulations (or through international agreements with neighboring countries) from levels
that are generally uncontrollable by the U.S.
Kim et al. (2014a) found increases in nitrate concentrations in the mixed layer of the
North Pacific Ocean, extending to near-shore areas off the west coast of the U.S., with
attendant changes in the status of N limitation. Because NH3 and NFU+ are so highly
soluble, they are likely to be removed in rain during ascent before trans-Pacific transport.
Because SO2 is much less soluble than NH3, it can be transported to the free troposphere
by the warm conveyor belt system before it is oxidized to SO42 in cloud droplets or on
the surfaces of mineral dust particles. The survivability of nitrate is intermediate,
depending on the form the nitrate takes.
Zhang et al. (2012a) computed N deposition rates from background sources and from
domestic anthropogenic sources using GEOS-Chem. According to their estimates, most
of the eastern U.S. and parts of states along the Pacific Coast received >10 kg/ha/yr N
deposition. The version of GEOS-Chem used (8.2.3) is the same as described in Zhang et
al. (201 la) and used in the 2013 ISA for Ozone and Other Photochemical Oxidants (U.S.
EPA. 2013c). Figure 2-43 shows contributions from domestic anthropogenic, foreign
anthropogenic, and natural emissions to total (wet + dry) annual nitrogen deposition over
the CONUS for 2006 calculated by Zhang et al. (2012a) using the GEOS-Chem global
scale CTM with a horizontal resolution of 1/2° by 2/3°.
June 2018
2-101
DRAFT: Do Not Cite or Quote

-------
Natural
0.1 0.3 1 2 3 6 8 12 16 0 5 10 20 30 50 70 90 [%]
[kg N ha ' a ]
Source; Zhang et al. (2012a).
Figure 2-43 Contributions to oxidized and reduced nitrogen deposition from
U.S.: anthropogenic (top), foreign anthropogenic (middle), and
natural sources (bottom).
Nitrogen deposition enhancement	Percentage contribution
US
anthropogenic
Foreign
anthropogenic
1	The upper panel of Figure 2-43 shows that highest values from U.S. anthropogenic
2	sources are found in the eastern U.S., in and downwind of the Ohio River Valley, and in
3	and around urban areas. The middle panel of Figure 2-43 shows highest contributions
4	from foreign anthropogenic sources in regions of the CONUS bordering Canada and
5	Mexico. Note the band of highest contribution in upper New York State as a result of
6	emissions in southern Canada. There is also some indication in the Pacific Northwest of
7	smaller contributions due to transport from Eurasia. The pattern of N deposition in the
8	simulation for natural sources in Figure 2-43 (bottom panel), however, shows maximum
June 2018
2-102
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
deposition throughout the central U.S., with the highest values over the Midwest,
reflecting a combination of NOy emissions from lightning in the south-central U.S., from
biomass burning throughout the Southeast, and from soils, mainly in the Midwest. The
background contribution to N deposition is typically <30% over the eastern U.S. and
typically 30 to 50% in the western U.S. where N deposition is already lower. Overall,
according to these simulations, U.S. anthropogenic emissions account for 78% of Nr
deposition over the CONUS. Foreign anthropogenic emissions and natural emissions
account for 6 and 16% respectively of total N deposition in this model simulation.
Background concentrations of SO2 were calculated using the MOZART-2 global model
of tropospheric chemistry (Horowitz et al.. 2003) and were presented in the 2008 ISA for
Sulfur Oxides (U.S. EPA. 2008c). Background SO2 concentrations are estimated to be
below 10 parts per trillion (ppt) over much of the U.S. Maximum background
concentrations of SO2 of -30 ppt are found in the western U.S. In the Northwest, where
there are large geothermal sources of SO2, the contribution of background sources to total
SO2 is 70 to 80%; however, absolute SO2 concentrations are still on the order of ~2 ppb
or less. With the exception of the West Coast, where volcanic SO2 emissions cause high
background concentrations, background sources contribute 20% because of the geothermal
sources of SO2 and oxidation of DMS in surface water of the eastern Pacific. In
comparison, values observed at several relatively remote sites cited in the 2008 ISA for
Sulfur Oxides (U.S. EPA. 2008c') ranged from 20 to 40 ppt.
As noted earlier, volcanic sources of SO2 in the U.S. are found in the Pacific Northwest,
Alaska, and Hawaii. The greatest potential domestic effects from volcanic SO2 occur on
the island of Hawaii. Nearly continuous venting of SO2 from Mauna Loa and Kilauea
produces SO2 in high concentrations of ~5 ppm lasting for periods of up to 1 hour [see
Figure 2-34 and Figure 2-35 in the 2008 ISA for Sulfur Oxides; (U.S. EPA. 2008cYI at
two national park sites near the Kilauea caldera and the nearby east rift zone. The latter
emits several times as much SO2 as the Kilauea caldera. The two measurement sites
within the national park are <3 km from the summit emission source and ~10 km from
the east rift source and are affected by the two sources during southerly and easterly
winds. A number of communities and population centers are within the same distance
from the east rift gas source that affects these two monitoring sites. When the normal
trade wind flows are disrupted, emissions from the sources can be brought directly to
these various communities. Because these communities are located at a similar distance
from the large east rift emission source as the national nark monitoring stations, it is
2-103
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
probable that these communities are subjected to SO2 concentrations as high as those
measured within Hawaii Volcanoes National Park.
When considering emission sources further afield, intercontinental transport of O3 and
PM has been the focus of efforts by the Task Force on Hemispheric Transport of Air
Pollution (HTAP). To the extent that N and S species are transported along with O3 and
PM, they also contribute to deposition following subsidence to the surface. Modeling
studies estimate that only a small fraction of nitrogen and sulfur emissions are transported
to and deposited within a continent different than the source of the emissions (Stock et
al.. 2013; Roy et al.. 2012; Sanderson et al.. 2008). Global scale modeling results
reported in the 2008 ISA (U.S. EPA. 2008a') and in the latest ISAs for Oxides of Nitrogen
Health Criteria (U.S. EPA. 2016f) and Sulfur Oxides Health Criteria (U.S. EPA. 2017c')
also indicate that intercontinental transport of oxidized and reduced nitrogen, SO2, and
S042 are likely minor background sources of these species. Of greater importance are
localized emissions from natural sources. These include emissions of NO from soils and
lightning and emissions of SO2 from geothermal and biogenic sources.
Background PM2 5 concentrations that would occur in the U.S. in the absence of
anthropogenic emissions in continental North America were reported in the 2009 PM ISA
(U.S. EPA. 2009a'). They include contributions from natural sources throughout the world
and from anthropogenic sources outside continental North America. These are both
primary and secondary natural and anthropogenic components. Estimated PM2 5 were
estimated to be less than 1 |ig/m3 on an annual basis, with maximum daily average values
in a range from 3 to 20 |ig/m3 and a peak of 63 |ig/m3 at the nine national park sites
across the U.S. (U.S. EPA. 2009a').
Just as pollutants can be transported into the U.S., they can also be transported outward.
For example, wet deposition of pollutants emitted in eastern North America occurs over
the North Atlantic Ocean. Deposition of N species is an important source of nutrients to
the western North Atlantic (Zamora et al.. 2011). Dennis et al. (2013). based on CMAQ
modeling results, estimated that -1/3 of oxidized N emissions and slightly less than 1/3
of NH3 emissions in the U.S. are transported out over the North Atlantic Ocean. Although
the average pH of rainwater at Bermuda is ~5, reflecting deposition of acidic species
emitted in North America, this additional source of acidity is at most only -2% of that
due to anthropogenic CO2 (Bates and Peters. 2007).
With well-validated models, it is possible to compare the relative role of different
emission and removal processes. Adj oint models (Henze et al.. 2009) are particularly
useful for understanding the relative contribution of emission sources to dry and wet
deposition of different nitrogen and sulfur containing compounds. For example, Lee et al.
(2016) found that half of nitrogen deposition at Federal Class I areas, such as national
2-104
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
parks, can be attributed to emission sources within 500 km and 90% of nitrogen
deposition is due to emission sources within 1,500 km. (Malm et al.. 2013) simulated
conservative tracer transport from ammonia source regions with the GEOS-Chem model
to estimate that roughly equal amounts of ammonia deposition in Rocky Mountain
National Park (RMNP), CO was from within or outside Colorado, with most of the
transport into Colorado coming from the West. Thompson et al. (2015) reported that 40%
reduced nitrogen deposition in RNMP was from outside Colorado.
Background rainwater pH and background deposition in remote areas worldwide has
considerably lower H+ and N deposition levels than in more populated areas, as described
in Appendix 2.6. Galloway et al. (1982) measured the pH of rainwater at five remote sites
worldwide and measured pH values ranging from 4.8 to 5.0. At some sites acidity was
attributed to long range transport of acid sulfate, while at others a mixture of strong and
weak acids attributed to both anthropogenic and natural sources was observed. They
concluded that a pH of 5 was a good lower limit estimate for natural contributions. Curtis
et al. (2018) estimated nitrogen and sulfur deposition in remote inland areas of Greenland
and reported 0.13 to 0.19 kg N/ha/yr for total nitrogen, 0.8 to 0.11 kg N/ha/yr for NOs .
0.05 to 0.09 kg N/ha/yr for NH4+, and 0.08 to 0.13 kg S/ha yr for SO42 .
2.7. Supplemental Material on Changes in Deposition since 2000
Maps on the portion of the NADP website dedicated to the Total Deposition (TDEP)
program (http://nadp. sws .uiuc.edu/committees/tdep/tdepmaps/) and shown in this section
present a comprehensive overview of changes in various parameters related to deposition
over 2000-2013. Changes between two, 3-year periods, 2000 through 2002 and 2011
through 2013 are summarized in this section. The most notable changes in the
geographical distribution of total (wet + dry) deposition of NOy + NHx between these
two time periods are found in areas with extensive agriculture (e.g., the San Joaquin
Valley, southern Idaho/northern Utah, and the Midwest), which show large increases in
total deposition of N (see Figure 2-44). There are also shifts in the distribution of wet and
dry deposition towards a greater predominance of dry deposition in these areas (see
Figure 2-45. Figure 2-46. and Figure 2-47). Deposition of oxidized nitrogen has declined
markedly throughout the eastern U.S. and southern California between the periods
2000-2002 and 2011-2013 (see Figure 2-48. Figure 2-49. Figure 2-50. and Figure 2-51)
due mainly to large decreases in dry deposition of total nitrate (TNO3 = HNO3 + pNOs :
see Figure 2-52). The decreases in total nitrate across the CONUS are generally due to
decreases in HNO3 (see Figure 2-53). Decreases in dry deposition of pNC>3- have
generally been smaller, reflecting the smaller contribution of pNC>3- to TNO3. Exceptions
include areas such as Florida, Texas, and southern California where dry deposition of
June 2018
2-105
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
pNCh- has been much greater than HNO3 in the earlier period and has decreased
substantially (see Figure 2-54). Other N species, mainly NO2 also show large decreases
(see Figure 2-55 and Figure 2-56). especially near urban source areas. Using OMI data,
Krotkov et al. (2016) found decreases in column (vertically integrated) abundances of
NO2 of -40% from 2005 to 2014 over the U.S.
In contrast to deposition of oxidized N, deposition of reduced inorganic N has seen large
increases (see Figure 2-57 and Figure 2-58). Several hotspots (the San Joaquin Valley,
southern Idaho, and several areas in the central and eastern U.S.) have increased
markedly in size (see Figure 2-57) between the two periods. Dry deposition of NH3 has
been the major contributor to the increase (see Figure 2-59). while dry deposition of
pNH4+ has largely decreased between the two periods (see Figure 2-60). Between the two
periods, emissions ofNOxhave decreased resulting in lower formation rates of HNO3
that could react with NH3 to form PNH4NO3.
Deposition of oxidized and reduced N have undergone geographic shifts with
corresponding shifts in the contributions of each to total N deposition on regional and
smaller scales (compare Figure 2-48 and Figure 2-49 to Figure 2-57 and Figure 2-58).
For example, deposition of oxidized N in the Northeast has decreased substantially, but
deposition of reduced N has increased substantially in the central U.S. Sizable shifts are
also seen in the fractional contributions of total N deposition as dry deposition of both
oxidized and reduced forms (compare Figure 2-47 and Figure 2-62).
Substantial declines in the deposition of S have occurred over the past 15 years,
particularly in the Ohio River Valley (see Figure 2-63). with generally much smaller
declines in wet deposition (see Figure 2-64) than for dry deposition (see Figure 2-65).
Despite these declines, S deposition is still highest in the Ohio River Valley. Except for
small shifts in several areas, the proportion of S dry deposited has been rather similar (see
Figure 2-66). Dry deposition of SO2 is still dominant over pS042—in eastern Ohio and
central Pennsylvania (compare Figure 2-66 and Figure 2-67 to Figure 2-68). These
decreases are consistent with those derived by Krotkov et al. (2016) who detected
decreases in the column (vertically integrated) abundance of SO2 of -75 % for the period
2005 to 2014 over the Ohio River Valley and southwestern Pennsylvania. These
decreases reflect reductions in emissions mandated by the Clean Air Act Amendments
and other regulatory requirements.
As noted earlier, these estimates of change in deposition were derived from CMAQ
output and data from measurement networks. Each of these components has its own set
of uncertainties, and the estimates of deposition and the changes over time should be
viewed in this light.
2-106
DRAFT: Do Not Cite or Quote

-------
Source: CAS'I'MmyCMAQ/NTO/AMON/SHARai
Total N
(kg-W/ha)
(:
-6
Total deposition of nitrogen 0002
USI2PA 1(1/16/1-1
sotitrft: c:astn fT/c:MAQ/NO"N/AM< )n/si:arci I
Total N
(kg-N/ha)
-6
n-Q
-10
J-12
1-14
::
>20
Total deposition of nitrogen 1113
uskpa 10/16/11
N = nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-44 Wet plus dry deposition of total nitrogen over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-107
DRAFT: Do Not Cite or Quote

-------
Wet N deposition 0002
USEPA 10/16/14
Source: CASTNhT/CMAQ/NTN/AM(JN/SEARCH
WetM
(kg-NJha)
Smu: CASTNrn/CMAQ/NTN/AMC)N/SIiAIiC:i i
Wet N deposition 1113
IPSKI'A 1(1/16/11
WetN
(kg-N/hal
N = nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-45 Wet deposition of total nitrogen over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-108
DRAFT: Do Not Cite or Quote

-------
Dry N
(kg-N/ha)
[!
-5
-6
-7
-8
Souire: CASTNiriYCIMAQ/NTN/AMC )N/S! 12
Dry N
(kg-N/ha)
I
-4
N = nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-46 Dry deposition of total nitrogen over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-109
DRAFT: Do Not Cite or Quote

-------
Dry M
(Pet o! Total)
0
1
L10

-20

-30

-40

-50

-60

-70

-80

-90
1
L>100
Sourr ft: CASTNI ri'/C :'M AQ/NTN / AM QN/SHARCII
Pet of total N as dry deposition 0002
US EPA KV16/II
Pet of tot al N as dry deposition 1113
Sourr ft: CAST NI iT7CMAQ/N*l "N/AM ON/SI iARCM	USEPA 10/16/1-1
N = nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-47 Percent of total nitrogen as dry deposition over 3-year periods.
Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-110
DRAFT: Do Not Cite or Quote

-------
Source: CASTNIrI'/CMAQ/NTN/AM( >N/S! AkCI I
Total deposition of oxidized N 0002
IISBPA 10/16/14
Total oxN
(kg-N/ha)
Total deposition of oxidized N 1113
USBPA 10/1 fi/M
Soureft: C:ASTNinVC:MAQ/NTN/AMON/Sl 2ARO I
Total oxN
(kg-N/ha)
oxN = oxidized nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-48 Wet plus dry deposition of oxidized nitrogen over 3-year periods.
Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-111
DRAFT: Do Not Cite or Quote

-------
Source.: C:ASTN 1 HVCM AQ/NTN/AM(>N/SI lARC.'l 1
Pet of total N as oxidized N 0002
IJSHPA 1 (VI6/14
-60
[70
eo
90
>100
Total oxN
(Pet of Total)
Pet of total N as oxidized N 1113
USKPA 10/16/1-1
Souxce:CASTOfcT/t'MAQ/KrN/AMON/SHAKCH
-50
-60
[70
80
90
>100
Total oxN
(Pet of Total)
oxN = oxidized nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-49 Percent of total nitrogen as oxidized nitrogen over 3-year periods.
Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-112
DRAFT: Do Not Cite or Quote

-------
Source: CASTMTIVCMAQ/NTN/AMON/SJ-ARClI
Dry deposition of oxidized N 0002
USIiPA 10/16/M
Sourcft: ( :astn i rr/c :m aq/ntn/amc jn/si jarci i
Dry deposition of oxidized N 1113
IJ SI'PA 1(V16/1-1
Dry oxN
(kg-N/ha)
oxN = oxidized nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-50 Dry deposition of oxidized nitrogen over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-113
DRAFT: Do Not Cite or Quote

-------
Dry oxN
(Pet of Total)
|
'-20
-30
-40
50
60
70
80
90
>100
Pet, of total N as dry oxidized N 0002
USEPA 10/16/1-1
Sonice-.CASl'Mm'/CMAQ/NTO/AMON/SEARCH
Pet of total N as dry oxidi zed N 1113
USEPA 10/16/1-1
t- 60
70
80
90
>100
Soutce: CASTN ET/CMAQ/NTN/AMUN/SEARCB
Dry oxN
(Pet of Total)
(»
-20
-30
-40
oxN = oxidized nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-51 Percent of total nitrogen dry deposited as oxidized nitrogen over
3-year periods. Top: 2000-2002; Bottom: 2011-2013,
June 2018
2-114
DRAFT: Do Not Cite or Quote

-------
Dry deposition of HN03 + pN03 0002
Source: CASTNFr/CMAQ/NTN/AMON/SBARCH	USEPA 10/16/14
-3.0
-4.0
y-5.o
B->6.0
Souie e: CAS'l'Nm /CMAQANTN /AMON/SliARCIl
Dry deposition of HN03 + pN03 1113
USEPA 10/16/1-1
TN03
(kg-N/ha)
--0.0
-2.0
TNO3 = nitric acid arid particulate nitrate.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-52 Combined dry deposition of nitric acid and particulate nitrate over
3-year periods. Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-115
DRAFT: Do Not Cite or Quote

-------
Dry deposition of nitric acid 0002
Source: CASTN ET/CM AQ/NTN/AM ON/SEAR CH	USEPA 10/16/14
Dry deposition of nitric acid 1113
USEPA 10/16/14
-4.0
|o
Source: CASTNET/CMAQ/NTN/AMON/SEARCH
HNO3
(kg-N/ha)
HNO3 = nitric acid.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-53 Dry deposition of nitric acid over 3-year periods. Top: 2000-2002;
Bottom: 2011-2013.
June 2018	2-116	DRAFT: Do Not Cite or Quote

-------
Dry deposition of particle nitrate 0002
USEPA 10/16/14
Some e: CASTN tfl/CMAQ/NlTS /AMOW/SBARCH
-1.4
f.
¦->2.0
pN03
(kg-N/ha)
r i
-0.0

-0.2

-0.4

-0.6

-0.8
-1.0
¦ 1.2
Source: CASTN I 'i'/OM AQ/NTN/AM( )N/SI -ARC! I
Dry deposition of particle nitrate 1113
USEPA 10/16/11
-1.0
-1.2
pN03 = particulate nitrate.
Source: CAST NET/CM AQ/NT N/AM ON/SEARCH.
Figure 2-54 Dry deposition of particulate nitrate over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-117
DRAFT: Do Not Cite or Quote

-------
Souire: CASUS! FvIVC:M AQ/N'l'N/AMC)N/S1 ARCH
-2.5
-3.0
t-3.5
4.0
4.5
>5.0
Dry deposition of unmeasured N species 0002
IJSEPA icy 16/1-1
o
rtf.
Other N
(kg-N/ha)
Souice: CAS1T4 lfl'/CM AQ/N'1'N /AM UN /SEARCH
Dry deposition of unmeasured N species 1113
USIII'A 10*16/1-1
-3.0
-3.5
-4.0
1-4.5
¦->5.0
Other N
(kg-N/ha)
E0.0
0.5
1°
1.5
-2.0
-2.5
N = nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-55 Dry deposition of modeled (unmeasured) nitrogen species over
3-year periods. Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-118
DRAFT: Do Not Cite or Quote

-------
Pet of total N as unmeasured species 0002
USKPA 10/16/M
Soiurft: CASTN Iri'/CMAQ/NTN/AM()N/SI!ARCI I
Sourrft: CASTN KM !M AQ/NTN/AM< )N/SKARC1 I
Pet of total N as unmeasured species 1113
USKPA 1 (VI6/11
Other N
(Pet of Total)
N = nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-56 Percent of total nitrogen as modeled (unmeasured) species over
3-year periods. Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-119
DRAFT: Do Not Cite or Quote

-------
Somen: CASTNBT/CMAQMTM/AMON/SBARa i
Total deposition of reduced N 0002
USEPA UVl&l 'l
Total reN
(kg-N/ha)
Total deposition of reduced N 1113
USEPA
Kniirca: t:AK'F'N[-:'[/C:'vl AQ/N'[7\7AM()Nffl[-:Ai«:il
reN = reduced nitrogen.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-57 Wet plus dry deposition of reduced (inorganic) nitrogen over
3-year periods. Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-120
DRAFT: Do Not Cite or Quote

-------
W / ^ \? i
Source: CASTNFT/CMAQ/NTN/AMONVSr.ARCII
Pet of total N as reduced N 0002
USEPA 10/16/1-1
-so
[70
SO
90
>100
Total reN
(Pet of Total)
Total reN
(Pa of Total)
[-
-20
""30
-40
-50
-60
I70
80
90
>100
Pet of total N as reduced N 1113
IJSHPA 10/16/14
Source: CASTNET/CM AQ/NTN/AMQN/SEARCH
reN = reduced nitrogen.
Source: CASTNET/CMAQ/NT N/AM ON/SEARCH.
Figure 2-58 Percent of total nitrogen deposition by reduced (inorganic)
nitrogen over 3-year periods. Top: 2000-2002: Bottom:
2011-2013.
June 2018
2-121
DRAFT: Do Not Cite or Quote

-------
Soiurft: CASTNBl'/C'MAQ/NTN/AMC)N/SFAROI
Dry deposition of ammonia 0002
USHPA l(VI6/1-1
NH3
(kg-N/ha)
-0
Soiure: C.'ASTNKiTCZMAQ^rm/AMDN/SEARCII
NH3
(kg-N/ha)
I
Dry deposition of ammonia 1113
USI'J'A l(Vlfi/14
NH3 = ammonia.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-59 Dry deposition of ammonia over 3-year periods. Top: 2000-2002;
Bottom: 2011-2013.
June 2018
2-122
DRAFT: Do Not Cite or Quote

-------
Soma: tlAST'Nlf!iCM Aq/NTN/AM< )N/SI-ARCII
Eiry deposition of particle ammonium 0002
USEPA lwiti/l 'l
PNH4
(kg-N/ha)
fo.o
0.2
0.4
0.6
-0.8
-1.0
-1.2
Some*: CASTNET/CMAQ/NTN/AMON/SUAKCH
Dry deposition of particle ammonium 1113
USEPA lonfi/ti
pNH4
(kg-Nrtia)
¦
-0.0

-0.2

-0.4

-0.6

-0.8
pNH4 = particulate ammonium.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-60 Dry deposition of particulate ammonium over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018	2-123	DRAFT: Do Not Cite or Quote

-------
fC*P\$ (
Source: CASTNET/CMAQ/NTN/AMON/SEARCH
Dry deposition of reduced N 0002
USEPA 10/16/14
Dry reN
(kg-N/ha)
fi
-4
Source: CAS'I'NiilVCMAQ/NTN/AMON/SliARCIi
Dry deposition of reduced N 1113
LI SliPA 10/16/11
Dry reN
(kg-N/ha)
reN = reduced nitrogen.
Source: CASTNET/CMAQ/NT N/AM ON/SEARCH.
Figure 2-61 Dry deposition of reduced (inorganic) nitrogen over 3-year
periods. Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-124
DRAFT: Do Not Cite or Quote

-------
Pet of total N as dry reduced N 0002
USEPA 10/16/14
Dry reN
(Pet of Total)
¦
r°

-10

-20

- 30

-40

-50

-60

-70

-80

^90
1
Lj-100
Source: CASWmyC'MAQ/NTN/AMON/SiiARC'Il
Pet of total N as dry reduced N 1113
USEPA 10/16/11
Dry reN
(Pet ol Total)
¦
r°

-10

-20

-30

-40

-50

-60

-70
I
-80

L90
¦
L>ioo
Souic e: CASTNEl'/CMAQ/JMl'N /AMON/SEARCI1
reN = reduced nitrogen.
Source: CASTNET/CMAQ/NT N/AM ON/SEARCH.
Figure 2-62 Percent of total nitrogen deposition by dry reduced (inorganic)
nitrogen over 3-year periods. Top: 2000-2002; Bottom:
2011-2013.
June 2018
2-125
DRAFT: Do Not Cite or Quote

-------
Soiirc#: CZASlWn/CMAQ/WrN/AMON/SEARCIl
Total deposition of sulfur 0002
11 SI {PA 10/16/14
Total S
(kg-S/ha)
Source: CASTNin/CMAQ/NTN/AMON/SliARCII
Total deposition of sulfur 1113
tlSUPA 10/16/1-1
Total S
(kg-S/ha)
S = sulfur.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-63 Wet plus dry deposition of total sulfur over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-126
DRAFT: Do Not Cite or Quote

-------
Soiurft: CASrI*NJnVCMAQ/NTN/AMON/SEARc:i I
Wet S deposition 0002
USI-I'A KVlti/M
WetS
(kg-S/ha)
Wet S deposition 1113
Source: CASTNHT/CMAQ/NTN/AMON/SHARCH	USEPA 10/16/14
S = sulfur.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-64 Wet deposition of total sulfur over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018	2-127	DRAFT: Do Not Cite or Quote

-------
Sane*: CASTN ITrt.'MAQWI'N/AMtlN/SIiARCH
Dry S deposition 0002
tlSEl'A 10/16/M
Dry S
(kg-SAia)
Sonms: CASTNinxrMAQftnWAMONjSliARCI)
DryS deposition 1113
USKPA 1 (VI6/1-1
Dry S
(kg-S/fia)
S = sulfur.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-65 Dry deposition of total sulfur over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-128
DRAFT: Do Not Cite or Quote

-------
Soiiirft: CASTN1 riyCMAQ/NTN/AMC )N/SI 2ARC1 I
-60
170
80
90
>100
Pet of total S as dry deposition 0002
USFiPA 10/16/M
DryS
(Pel of Total)
¦
-0
P
-10

-20

-30

-40
Soiurft: CA5TNI -T/CIM AQ/N*I"N/AM ON/SEARCl I
Pet of total S as dr>r dqposition 1113
USBPA 10/16/M
Dry S
(Pet of TotaJ)
I
-20
-30
-40
S = sulfur.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-66 Percent of total sulfur deposition by dry deposition over 3-year
periods. Top: 2000-2002; Bottom: 2011-2013.
June 2018
2-129
DRAFT: Do Not Cite or Quote

-------
Soiurft: CASTN fTiCM AQ/NTN/AM( )N/SHARCH
Dry deposit ion of sulfur dioxide 0002
IJSEPA 1(V1(S/14
S02
(kg-S/ha)
[:
-B
Drjr deposition of sulfur dioxide 1113
USEPA 10/16/14
Soturft: CASTN mVCMAQ/NTN/AMf>N/SEAKC.I I
S02
(kg-S/ha)
¦
-0

-2

-4

-6

— 8

-10

-12

-14

-15

- 1B

-20

-22

->24
S02 = sulfur dioxide.
Source: CAST NET/CMAQ/NT N/AM ON/SEARCH.
Figure 2-67 Dry deposition of sulfur dioxide over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018
2-130
DRAFT: Do Not Cite or Quote

-------
Soutee,: CASTNET/CM AQ/NTN/AM ON/SFARCH
P
¦->2.0
Dry deposition of particle sulfate 0002
USEPA 10/16/14
pS04
(kg-S/ha)
¦
-0.0

-0.2

-0.4
•
-0.6

-0.8
Dry deposition of particle sulfate 1113
USEPA 10/10/14
Source: CASTNET/CMAQ/NTN/AMON/SEARCH
pS04
jkg-S/ha)

-0.0
1
-0.2

-0.4
¦
-0.6

-0.8
pS04 = particulate sulfate.
Source: CASTNET/CMAQ/NTN/AMON/SEARCH.
Figure 2-68 Dry deposition of particulate sulfate over 3-year periods. Top:
2000-2002; Bottom: 2011-2013.
June 2018	2-131	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
APPENDIX 3. DIRECT PHYTOTOXIC EFFECTS OF
GASEOUS OXIDIZED NITROGEN
AND SULFUR ON VEGETATION
This appendix provides a brief overview of the exposure and phytotoxic effects of
gaseous forms of oxidized nitrogen (N) and sulfur (S) compounds on vegetation. The
main focus of this ISA is understanding the ecological impact of oxidized N and S and
that the major effects of these compounds on ecosystems are through acidifying
deposition and N enrichment deposition. However, direct effects of gaseous oxidized N
and S could augment the effects of deposition on vegetation, and direct effects of gaseous
N and S may be apparent in some areas. The effect of sulfur dioxide (SO2) gas on
vegetation is discussed in Appendix 3_2. Appendix 33 discusses the effects of nitric
oxide (NO), nitrogen dioxide (NO2), and peroxyacetyl nitrate (PAN) on vegetation.
Appendix 3.4. presents information on the direct effects of nitric acid (HNO3) vapor on
vegetation, including lichens. A summary section with causal determinations based on a
synthesis of the body of information on the biological effects of exposure to these gases
is presented in Appendix 3.5.
3.1. Introduction
The effects of gaseous pollutants such as SO2, NO2, NO, HNO3, and ozone (O3) on
vegetation have been studied since at least the early 19th century (Holmes etal.. 1915;
Havwood. 1905). Methodologies have been developed to study the effects of gaseous
exposures to these pollutants in the laboratory, greenhouse, and in the field. The
methodologies to study the effects of gaseous pollutants on vegetation have been recently
reviewed in the 2013 Ozone ISA and 2006 Ozone AQCD [Air Quality Criteria
Document; (U.S. EPA. 2013c. 2006a)]. A thorough description of the methodologies
used to expose vegetation to gaseous pollutants can be found in Section 9.2 of the 2013
Ozone ISA (U.S. EPA. 2013c). AX9.1 of the 2006 Ozone AQCD (U.S. EPA. 2006a). and
Section 9.2 in the 1993 Oxides of Nitrogen AQCD (U.S. EPA. 1993).
Uptake of gaseous pollutants in a vascular plant canopy is a complex process involving
adsorption to surfaces (leaves, stems, and soil) and absorption into leaves. These
pollutants penetrate into leaves primarily in gaseous forms through the stomata. The
surface cuticle provides a protective barrier to gaseous pollutant exposure, although there
is evidence for limited uptake across the cuticle (Zhang et al.. 2003; Kerstiens et al..
1992). Pollutants must be transported from the bulk air to the leaf boundary layer to reach
the stomata. The transport of pollutants through a boundary layer into the stomatal region
June 2018
3-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
is by diffusion. Studies of transport through the boundary layer are based on aerodynamic
concepts and usually relate to smooth surfaces that are not typical of leaf-surface
morphology (Gates. 1968). Once through the boundary layer, the gas enters the leaf
through the stomata. The entry of gases into a leaf is dependent upon gas-phase chemical
processes and physical characteristics of surfaces, including stomatal aperture. The
aperture of the stomata is controlled largely by the prevailing environmental conditions,
such as humidity, temperature, light intensity, and water availability. When the stomata
are closed, as occurs under dark or drought conditions, resistance to gas uptake is very
high and the plant has a very low degree of susceptibility to injury (Figure 3-1). The
stomatal control of uptake of gaseous pollutants is described in more detail in AX9.2 of
the 2006 Ozone AQCD (U.S. EPA. 2006a) and Section 9.3.1.5 of the 1993 Oxides of
Nitrogen AQCD (U.S. EPA. 1993). Note that unlike vascular plants, mosses and lichens
do not have a protective cuticle barrier to gaseous pollutants, which is a major reason for
their sensitivity to gaseous S and N.
3-2
DRAFT: Do Not Cite or Quote

-------
Light
O3. NOx, SOx
Cuticle
Epidermis
Pallisade
Mesophyll
Spongy
Mesophyll
Epidermis
Cuticle
Vascular
-System
h2o
c0= [C02]
__ij
Guard Cell
03, NOx, SOx
Abbreviations: C, = internal C02 in leaf; C0 = C02 of the atmospheric air; C02 = carbon dioxide; H20 = water; SOx = sulfur oxides;
NOx = oxides of nitrogen; 03 = ozone.
Figure 3-1 The microarchitecture of a dicot leaf. While details among species
vary, the general overview remains the same. Light that drives
photosynthesis generally falls upon the upper (adaxial) leaf
surface. Carbon dioxide, oxides of sulfur, oxides of nitrogen, and
ozone gases generally enter by diffusion through the guard cells
(or stomata) on the lower (abaxial) leaf surface, while water vapor
exits through the stomata (transpiration).
3.2.	Direct Phytotoxic Effects of Sulfur Dioxide on Vegetation
1	It has been known since the early 1900s that exposure to SO2 can cause plant damage and
2	death (Wislicenus. 1914). The large sources of historic SO2 emissions were ore smelters.
3	Sulfides m the ore were oxidized during smelting and resulted in large releases of SO2.
4	Emissions from large ore smelters in the U.S. and Canada resulted in large areas denuded
5	of vegetation surrounding these facilities (Thomas. 1951; Swain. 1949). Much of the
June 2018
3-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
damage to the vegetation was due to acute effects of high concentrations of SO2.
However, as early as 1923, researchers recognized that SO2 might decrease plant growth
without producing acute symptoms of foliar injury (Stoklasa. 1923). In the 1950s through
the early 1980s, there was much research on the effects of SO2 on vegetation, as well as
the interaction with pollutants such as O3 and NO2. Since then, there has been much less
research on the effects of SO2 on vegetation, especially in the U.S., due to the decreasing
ambient concentrations of SO2 (U.S. EPA. 2012a) (see Figure 3-2 for max 3-hour SO2
concentrations for 2016). The effects of SO2 on vegetation are summarized below.
Currently, SO2 is the only criteria pollutant with a secondary NAAQS distinct from the
primary standard. This standard is to protect acute foliar injury resulting from SO2
exposure. The standard is a 3-hour average of 0.50 ppm and was promulgated in 1971 to
protect against the adverse effects of acute foliar injury in vegetation. The 1982 AQCD
for Particulate Matter and Sulfur Oxides concluded that controlled experiments and field
observations of vegetation supported retaining this secondary standard (U.S. EPA. 1982b.
d. 1971).
Acute foliar injury usually occurs within hours of exposure, involves a rapid absorption
of a toxic dose, and involves a collapse or necrosis of plant tissues. Another type of
visible injury is termed chronic injury and is usually a result of variable SO2 exposures
over the growing season. After entering the leaf, SO2 is converted to sulfite (SOr ) and
bisulfite (HSO3 ) ions, which may be oxidized to sulfate (SO42 ). Sulfate is about
30 times less toxic than sulfite and bisulfite. The conversion of sulfite and bisulfite to
sulfate results in net H+ production in the cells. Kropff (1991) proposed that the
appearance of SC>2-induced leaf injury was likely due to a disturbance of intracellular pH
regulation. Kropff (1991) listed several studies in which the pH of homogenates of leaf
cells only shifted towards greater acidity when plants were lethally damaged from
long-term SO2 exposures (Jager and Klein. 1977; Grill. 1971; Thomas et al.. 1944). The
appearance of foliar injury can vary significantly among species and growth conditions
(which affect stomatal conductance). Currently, there is no regular monitoring for SO2
foliar injury effects in the U.S.
3-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
Figure 3-2 Map of maximum 3-hour daily max average sulfur dioxide
concentration reported at Air Quality System monitoring sites for
2016.
Besides foliar injury, long-term lower SO2 concentrations can result in decreased
photosynthesis, growth, and yield of plants. These effects are cumulative over the
growing season and are often not associated with visible foliar injury. As with foliar
injury, the effects of these injuries vary among species and growing environment. The
1982 Particulate Matter and Oxides of Sulfur (PM-SOx) AQCD summarized the
concentration-response information available at the time (U.S. EPA. 1982b). Effects on
growth and yield of vegetation were associated with increased SO3 exposure
concentration and time of exposure. However, that document concluded that more
definitive concentration-response studies were needed before useable exposure metrics
could be identified. Because ambient SO2 concentrations declined and focus on O3
vegetation effects research increased, relatively few studies have emerged to better
June 2018
3-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
inform a metric and levels of concern for effects of SO2 on growth and productivity of
vegetation.
SO2 is considered to be the primary factor causing the death of lichens in many urban and
industrial areas, with fruticose lichens being more susceptible to SO2 than many foliose
and crustose species (Hutchinson etal.. 1996). Damage to lichens in response to SO2
exposure includes reduced photosynthesis and respiration, damage to the algal
component of the lichen, leakage of electrolytes, inhibition ofN fixation, reduced K+
absorption, and structural changes (Hutchinson et al.. 1996; Belnap et al.. 1993; Farmer et
al.. 1992). Significant reductions in lichen photosynthesis have been measured at
concentrations as low as 91 ppb over 2-4 hours (Sanz et al.. 1992; Huebert et al.. 1985).
Damage to the algal component of the thallus is evidenced by its discoloration. The entire
thallus dies soon after algal cells are damaged (Hutchinson et al.. 1996). At higher
concentrations, SO2 deactivates enzymes by chemical modification, leading to reduced
metabolic activity and loss of membrane integrity (Nieboer et al.. 1976; Ziegler. 1973). In
addition, SO2 binds to the central metal atoms of enzymes, adversely affecting membrane
function and cell osmolality. SO2 also competitively inhibits bicarbonate (HCO, ) and
dihydrogen phosphate (H2PO4 ) interactions with enzymes (Hutchinson et al.. 1996).
Low pH increases the toxicity of SO2 action (Farmer et al.. 1992). The toxic effects of
atmospheric deposition of SO2 are lessened when lichens are attached to a substrate,
typically bark or rock, that has high pH or superior buffering capacity (Richardson and
Cameron. 2004). van Herk (2001) evaluated relationships between bark pH and air
pollution levels as two significant variables affecting epiphytic lichen composition and
concluded that bark pH was the primary factor regulating the distribution of acidophilic
species in the Netherlands. In studies of unpolluted areas, differences in bark chemistry
also affect the presence and distribution of epiphytes (Farmer etal.. 1992). Indirect
changes to bark pH, caused by acidification and high SO2 concentrations, also affect
lichen distribution (Farmer et al.. 1992). In a more recent, Geiser and Neitlich (2007)
reported that direct SO2 damage to lichens in the Pacific Northwest may have been
confined to major urban areas such as Seattle, Portland, and Belingham. However, lichen
monitoring plots were not colocated with SO2 monitors and the authors were not able to
quantify SO2 exposure. More information on the N effects reported in this study is found
in Appendix 6.3.7 and Appendix 6.5.2.
More recent research has been performed in areas of Europe where ambient SO2
concentrations are generally higher than in the U.S. Since the 2008 ISA, several studies
in Germany and some eastern European countries have indicated that direct effects of
SO2 caused growth reductions in trees during the last century (Cavlovic et al.. 2015;
Hauck et al.. 2012; Rvdval and Wilson. 2012; Elling et al.. 2009). Elling et al. (2009)
evaluated a large database providing long-term growth records of 1,010 silver firs (Abies
3-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
alba), long-term climate records, and long-term air pollution data from 51 sites in
southern Germany. In this analysis, silver fir growth was influenced by SO2 pollution
more than any other factor in the second half of the last century. The study authors also
reported an almost immediate increase in growth in response to falling SO2 emissions in
the 1980s. This rapid response indicates a direct effect of gaseous SO2 rather than an
indirect effect of soil acidification, which would have a longer response time. Annual
average concentrations above 10 jj.g SCh/m3 (approximately 4 ppb) appeared to reduce
growth of silver fir. (See Figure 2-15 for recent concentrations in the continental U.S.). In
a later publication, Bosela et al. (2014) argued a combination of SO2 and NOx gas
emissions has historically reduced growth in silver fir in the western Carpathian
Mountains of the Czech Republic.
A similar gaseous SO2 effect on tree growth may have been occurring in the eastern U.S.
Using tree ring analysis, Thomas et al. (2013) reported significant growth increases in
old-growth eastern red cedar (Juniperus virginiana) in West Virginia following decreases
in SO2 emissions since 1980. Growth continued to increase as SO2 emissions further
declined in the 1990s and 2000s. Thomas et al. (2013) also found evidence of
physiological changes in response to SO2 emissions. The authors attributed the growth
response to an indirect effect of decreasing acidifying deposition, and thus, recovery from
soil acidification. However, a historical record of acidifying deposition was not available.
As in Europe, the trees studied in West Virginia also had a relatively rapid recovery in
response to declining SO2 emissions that could indicate the effects were from direct
exposure to gases in the atmosphere rather than soil acidification. Further, a response to
this study from other researchers suggested that the eastern red cedars in the West
Virginia study were found on a limestone outcrop that could be well buffered from soil
acidification (Schaberg et al.. 2014). This study may indicate that gaseous SO2 alone or in
combination with other gases may have inhibited red cedar growth. See Appendix 5.2.1.3
for further discussion of this study.
3.3. Direct Phytotoxic Effects of Nitric Oxide, Nitrogen Dioxide,
and Peroxyacetyl Nitrate
In sufficient concentrations, nitric oxide (NO) and nitrogen dioxide (NO2) can have
phytotoxic effects on plants by decreasing photosynthesis and inducing visible foliar
injury (U.S. EPA. 1993). The current secondary (welfare) and primary (human health)
standard for oxides of nitrogen is a NO2 annual mean of 0.053 ppm. See Figure 3-3 for
recent concentrations ofNC>2. The 1993 Oxides of Nitrogen AQCD concluded that
concentrations of NO2 or NO in the atmosphere are rarely high enough to have
phytotoxic effects on vegetation (U.S. EPA. 1993). Since the 1993 Oxides of Nitrogen
June 2018
3-7
DRAFT: Do Not Cite or Quote

-------
1	AQCD, very little new research has been done on these phytotoxic effects to alter this
2	conclusion (Bender and Weigel. 2011). However, it is known that these gases alter the N
3	cycle in some ecosystems, especially in the western U.S., and contribute to N saturation
4	(Sparks. 2009; Fenn et al.. 2003a: Bvtnerowicz and Fenn. 1996). See Appendix 6J_ for a
5	discussion of the nutrient effects of N.
Note: N02 = nitrogen dioxide. Concentrations indicated are the highest concentration in the county and do not represent countywide
concentrations.
Source: U.S. Environmental Protection Agency 2014 analysis of data from state and local air monitoring stations (U.S. EPA, 2016f).
Figure 3-3 Map of U.S. annual average nitrogen dioxide concentrations for
2013.
6	In general, NO and NO: enters leaves through the stomata (Saxe. 1986). However, the
7	leaf cuticle could be an important receptor for NO2, and there is evidence of transport of
8	NO and NO2 across isolated cuticles (Lendzian and Kerstiens. 1988). Several studies
9	have demonstrated that plant canopies can directly assimilate N in the form of NO2, but
June 2018
3-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
canopy uptake of NO2 is generally small relative to total plant uptake (Vallano and
Sparks. 2008; Ammann et al.. 1995; Nussbaum et al.. 1993; Segschneider et al.. 1993;
von Ballmoos et al.. 1993; Hanson et al.. 1989). After entering the leaves, NO2 dissolves
in the extracellular water of the substomatal cavity to form HNO2 and HNO3, which then
dissociate to form NO2 . NO3 . and H+ (Bvtnerowicz et al.. 1998). Both cell and tonoplast
membranes contain ATP-dependent H+ pumps, and the tonoplast pumps are strongly
inhibited by NO3 (Bvtnerowicz et al.. 1998). If extra protons are deposited in vacuoles of
the plant cells during normal cellular regulation, then additional acidity will occur in
combination with additional NO3 . This combination can cause disruptions in cellular
control (Taylor and MacLean. 1970). NO3 and nitrite (NO2 ) are metabolized to amino
acids and proteins through a series of enzymatic reactions mainly involving NO3 and
nitrite reductases (Amundson and MacLean. 1982). The ability of plants to reduce NO3
and NO2 to amino acids and proteins determines the potential of the plant to detoxify
NO and NO2 (Well burn. 1990). Reduction of NO3 takes place outside of the chloroplast
while the reduction of NCV is coupled with the light reactions of photosynthesis.
Therefore, when leaves are exposed to NO and NO2 in the dark, highly phytotoxic levels
of NO2 accumulate and may lead to greater toxicity to NO and NO2 at night (Amundson
and MacLean. 1982). Exposure to NO produces both NO3 and NO2 in the leaves, but
the rate of NO3 accumulation is much slower than NO2 . Thus, plants exposed to high
NO could be at risk to elevated concentrations of NO2 (Wcllburn. 1990). More detailed
information on the cellular effects of NO and NO2 can be found in the 1993 Oxides of
Nitrogen AQCD.
The functional relationship between ambient concentrations of NO or NO2 and a specific
plant response, such as foliar injury or growth, is complex. Factors such as inherent rates
of stomatal conductance and detoxification mechanisms and external factors, including
plant water status, light, temperature, humidity, and the particular pollutant exposure
regime, all affect the amount of a pollutant needed to cause symptoms of foliar injury.
Plant age and growing conditions and experimental exposure techniques also vary widely
among studies quantifying the response of plants to NO2. An analysis conducted in the
1993 Oxides of Nitrogen AQCD of over 50 peer-reviewed reports on the effects of NO2
on foliar injury indicated that plants are relatively resistant to NO2, especially compared
to foliar injury caused by exposure to O3 (U.S. EPA. 1993). With few exceptions, visible
injury has not been reported at concentrations below 0.20 ppm, and these exceptions
occurred when the cumulative duration of exposures extended to 100 hours or longer. At
0.25 ppm, increased leaf abscission was reported on navel orange trees (Citrus sinensis),
but only after exposures in excess of 1,000 hours (Thompson et al.. 1970). Green bean
(Phaseolus vulgaris) plants used as bioindicators of NO2 injury in Israel developed foliar
injury symptoms when ambient concentrations exceeded 0.5 ppm (Donagi and Goren.
1979). In most plants, injury occurred in less than 1 day only when concentrations
3-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
exceeded 1 ppm (U.S. EPA. 1993). In recent years (2011-2013), ambient hourly NO2
concentrations in the U.S. have been well below the exposures in the above studies with
maximum highest 1-hour daily concentrations less than 0.075 ppm [see Section 2.5.2 of
U.S. EPA (2016f)l.
Decreased rates of photosynthesis have been recorded in experimental exposures of
plants to both NO and NO2, but usually at concentrations significantly higher than would
normally be encountered in ambient air. For example, Sabaratnam et al. (1988) reported
that soybeans (Glycine max) exposed 7 hours/day for 5 days showed an increase in
photosynthesis at a concentration of 0.2 ppm but a decrease in net photosynthesis at a
concentration of 0.5 ppm. Short-term exposures of soybean to 0.6 ppm NO2 for 2 to
3 hours also had no effect on net photosynthesis (Carlson. 1983). Most plants appear to
be more susceptible to NO than to NO2, as shown by Saxe (1986). who exposed a variety
of horticultural plants raised in greenhouses (species of Hec/era. Ficus, Hibiscus,
Nephrolepis, and Dieffenbachia) to both NO and NO2. Saxe (1986) reported that
decreases in net photosynthesis occurred at doses of NO that were 22 times less than that
for NO2. However, these decreases in net photosynthesis required concentrations as high
as 1 ppm NO for 12 hours to elicit a response in these plants.
In the 1970s and 1980s, hundreds of studies were conducted on the effects of NO2 on
growth and yield of plants. These studies varied widely in plant species, growing
conditions, exposure equipment, concentrations, durations, exposure regimes, and
environmental conditions during exposures. No clear dose-response relationships for
exposure to NO2 and reductions in growth and/or yield of plants emerged from these
experiments. Readers are referred to the analysis of over 100 studies conducted in the
1993 Oxides of Nitrogen AQCD. A few key studies are highlighted in this section. The
growth of several plant species appears to be susceptible to concentrations of NO2 less
than 0.2 ppm, particularly when exposure occurs during low-light conditions. For
example, nearly continuous exposure to 0.1 ppm NO2 for 8 weeks significantly reduced
growth of Kentucky blue grass [Poa pratensis; (Whitmore and Mansfield. 1983;
Ashenden. 1979)1. Eight species of tree seedlings were exposed to 0.1 ppm NO2 for
6 hours/day for 28 days, resulting in reduced shoot or root growth in two species, white
ash (Fraxinus americana) and sweetgum (Liquidambar styraciflua), reduced height
growth in two clones of loblolly pine (Pinus taeda), and no effects on the other species
(Kress and Skellv. 1982). No effects of NO2 at 0.1 ppm or lower were observed on
numerous other species, including potato (Solarium tuberosum), black poplar (Populus
nigra), radish (Raphanus sativus), soybean, or peas (Pisum sativum) (U.S. EPA. 1993).
No effects of NO2 were observed on soybeans grown in field plots subjected to a series of
10 episodic exposures averaging 0.4 ppm for 2.5 or 3 hours (Irving et al.. 1982).
Numerous studies have reported negative effects on growth of a variety of plants exposed
3-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
to 0.5 ppm NO2 and above (U.S. EPA. 1993). but these concentrations are very high
relative to current ambient levels of NO2 (see Appendix 2.6.1).
The 1993 Oxides of Nitrogen AQCD reviewed the extensive literature on the effects of
NO2 in combination with other gaseous air pollutants, particularly SO2 and O3, and
concluded that combinations of pollutants can cause foliar injury or decreases in
photosynthesis at concentrations lower than those associated with NO2 acting alone.
However, the plant responses occur at concentrations much higher than typically found in
ambient air in the U.S. (U.S. EPA. 1993). In addition, the presence of NO2 in studies
combining other gases did not produce symptoms different from those caused by the
dominant pollutant, either SO2 or O3, such that a plant response produced by
combinations of NO2 with other air pollutants in the field would be difficult, if not
impossible, to distinguish from those of the other single pollutants (U.S. EPA. 1993).
Since the 1993 Oxides of Nitrogen AQCD was completed, most new research on NO2
exposure to vegetation has taken place in Europe and other areas outside the U.S. For
example, foliar NO3 reductase activity was increased in Norway spruce (Picea abies)
growing near a highway in Switzerland with average exposures of about 0.027 ppm
compared to trees growing 1,300 m away from the highway with NO2 exposures less than
0.005 ppm (Ammann et al.. 1995). This result was consistent with other studies on
Norway spruce in the field and laboratory (von Ballmoos et al.. 1993; Thoene et al..
1991). Muller et al. (1996) found that the uptake rate of NO3 by roots of Norway spruce
seedlings was decreased by exposure to 0.1 ppm ofN02 for 48 hours. Similarly, soybean
plants grown in Australia had decreased NO3 uptake by roots and reduced growth of
plants exposed to 1.1 ppm of NO2 for 7 days (Qiao and Murray. 1998). In a Swiss study,
poplar (Populus *euramericana) cuttings exposed to 0.1 ppm ofN02 for approximately
12 weeks resulted in decreased stomatal density and increased specific leaf weight, but
did not result in other effects such as leaf injury or a change in growth (Gi'inthardtgocrg et
al.. 1996). However, NO2 enhanced negative effects of ozone on poplars, including leaf
injury, when the pollutants were applied in combination (Gi'inthardtgocrg et al.. 1996).
Since the 2008 ISA for Oxides of Nitrogen and Oxides of Sulfur-Ecological Criteria
[hereafter referred to as the 2008 ISA (U.S. EPA. 2008a)l. very few studies have been
published on the direct effects of NO and NO2 on vegetation. Hu et al. (2015b) exposed
clonal hybrid poplar (Populus alba x Populus berolinensis) saplings to 4 ppm of NO2.
The authors reported significant declines in photosynthesis and dark respiration with
exposures of 48 hours. They also reported stomatal dysfunction at this level of exposure
resulting in partial stomatal closure and a decline in stomatal conductance. However,
4 ppm of NO2 is very high relative to current ambient levels of NO2 in the U.S. (see
3-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Appendix 2.6.1). These results are consistent with past studies of plants with relatively
high NO2 exposure.
As part of a broad study of N deposition of a native serpentine grassland in California,
Vallano et al. (2012) exposed five native grasses and forbs (Plantago erecta, Layia
gaillardioides, Lasthenia californica, Vulpia microstachys, and Cryptantha flaccida) and
the most common invasive grass Lolium multiflorum to NO2 and soil N addition. The
plants were exposed in a growth chamber to very low levels of NO2 (0.03 ppm) to
simulate recent ambient air concentrations for this ecosystem. At this relatively low NO2
exposure, no significant effects were found on shoot biomass, root biomass,
photosynthesis, or stomatal conductance. The authors reported that despite not finding
species responses to NO2 exposure, the additive effects of NO2 combined with soil N on
plant performance indicate that uptake of NO2 may play a role in species responses to
increasing N deposition. The authors found that the combined NO2 and N addition
resulted in a strong positive growth and competitive response in the invasive Lolium. This
result is consistent with previous findings that low levels of NO2 that are not phytotoxic
to plants can add to the N load to an ecosystem from uptake through leaves (Sparks.
2009).
In a study in the Grand Canyon National Park, AZ, Kenkel et al. (2016) found that NOx
concentrations measured by Ogawa passive samplers were about 52% higher along
roadsides than 30 m away from the road. The pattern of the amount of 15N in pinyon pine
(Pinus edulis) mirrored the concentration gradient of NOx concentrations from the road,
indicating that the vegetation is taking up N from the vehicle traffic in the park. The
authors reported that sustained chronic N deposition on this arid environment could result
in deleterious effects for these ecosystems.
Peroxyacetyl nitrate (PAN) is a well-known photochemical oxidant that often co-occurs
with O3 during high photochemical episodes and that has been shown to cause injury to
vegetation (See reviews by, Cape. 2003; Kleindienst. 1994; Temple and Taylor. 1983).
Acute foliar injury symptoms resulting from exposure to PAN are generally characterized
as a glazing, bronzing, or silvering of the underside of the leaf surface; some sensitive
plant species include spinach, Swiss chard, lettuces, and tomatoes (Temple and Taylor.
1983). Petunias (Petunia hybrida) have also been characterized as sensitive to PAN
exposures and have been used as bioindicators in areas of Japan (Nouchi etal.. 1984).
Controlled experiments have also shown significant negative effects on the net
photosynthesis and growth of petunias and kidney beans (Phaseolus vulgaris) after
exposure of 30 ppb of PAN for 4 hours on each of 3 alternate days (Cape. 2003; Izuta et
al.. 1993). As mentioned previously, it is known that oxides of N, including PAN, could
be altering the N cycle in some ecosystems, especially in the western U.S., and
3-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
contributing to N saturation (Fcnn et al.. 2003a; Bvtnerowicz and Fenn. 1996). Although
PAN continues to persist as an important component of photochemical pollutant
episodes, there is little evidence in recent years to suggest that PAN poses a significant
risk to vegetation in the U.S.
3.4. Direct Phytotoxic Effects of Nitric Acid
Relatively little is known about the direct effects of HNO3 vapor on vegetation. Recent
information on HNO3 concentrations are given in Appendix 2.6.3 The deposition velocity
of HNO3 is very high compared to other pollutants (see Table 2-2) and HNO3 may be an
important source of N for plants (Hanson and Garten. 1992; Hanson and Lindberg. 1991;
Yose and Swank. 1990). This deposition could contribute to N saturation of some
ecosystems close to sources of photochemical smog (Fenn et al.. 1998). For example, in
mixed conifer forests of the Los Angeles basin mountain ranges, HNO3 has been
estimated to provide 60% of all dry deposited N (Bvtnerowicz et al.. 1998). Since the
2008 ISA, a controlled exposure study (Padgett et al.. 2009a) reported that 10 to 60% of
the HNO3 retained by foliage was incorporated into the biologically active N pool. The
remainder of the HNO3 was bound to foliar surfaces. This new study provides further
evidence for HNO3 as a contributor of biologically available N to southern California
forests.
Norbv et al. (1989) reported that exposure of 75 ppb of HNO3 for 1 day increased nitrate
reductase activity in red spruce (Picea rubens) foliage. In another study, foliar nitrate
reductase activity was also increased in California black oak (Quercus kelloggi), canyon
live oak (Quercus chrysolepis), and ponderosa pine (Pinus ponderosa) seedlings with
exposure to 65 to 80 ppb of HNO3 for 24 hours (Krvwult and Bvtnerowicz. 1997).
Because the induction of nitrate reductase activity is a step in a process leading to the
formation of organic N compounds (amino acids), the nitrate from HNO3 could function
as an alternative source of N for vegetation (Calanni etal.. 1999). However, in plants
under stress, the reduction of nitrate to amino acids consumes energy needed for other
metabolic processes.
At high ambient concentrations, HNO3 can cause vegetation damage. Seedlings of
ponderosa pine and California black oak subjected to short-term exposures from
50-250 ppb of HNO3 vapor for 12 hours showed deterioration of pine needle cuticle at
50 ppb in light (Bvtnerowicz et al.. 1998). Oak leaves, however, appeared to be more
resistant to HNO3 vapor, with 12-hour exposures in the dark at 200 ppb producing
damage to the epicuticular wax structure (Bvtnerowicz et al.. 1998). The observed
changes in wax chemistry caused by HNO3 and accompanying injury to the leaf cuticle
June 2018
3-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
(Bvtnerowicz et al.. 1998) may predispose plants to various environmental stresses such
as drought, pathogens, and other air pollutants. Because elevated concentrations of HNO3
and ozone co-occur in photochemical smog (Solomon et al.. 1988). synergistic
interactions between the two pollutants are possible (Bvtnerowicz et al.. 1998). Note,
however, that the experiments described above were observed at relatively short-term
exposures at above ambient concentrations of HNO3. Studies of long-term effects of
lower air concentrations that more closely approximate ambient HNO3 are needed.
Since the 2008 ISA, Padgett et al. (2009b) investigated dry deposition of HNO3 on the
foliage of ponderosa pine, white fir (Abies concolor), California black oak, and canyon
live oak in southern California. Using a chamber system within a greenhouse, leaves and
needles were exposed to control (0 (.ig/nr1 HNO3), moderate (30 to 60 (.ig/nr1 peak HNO3),
and high (95 to 160 (ig/m3 peak HNO3) concentrations. The high concentrations
represented a high ambient concentration that occurs periodically in California. The
experimental exposures resulted in a suite of damage symptoms that intensified with
increasing exposure. The exposures caused substantial perturbations to the epicuticular
surfaces of foliage of all four tree species studied. The damage caused by dry deposition
may leave foliage more vulnerable to other copollutants such as ozone.
It has been suspected that HNO3 may have caused a dramatic decline in lichen species in
the Los Angeles basin (Nash and Sigal. 1999). The suggestion was strengthened by
transplant of Ramalina lichen species from clean air habitats (Mount Palomar and San
Nicolas Island) to analogous polluted habitats in the Los Angeles basin and observing
death of the lichens over a few weeks in the summer (Boonpragob and Nash. 1991).
Associated with this death was a massive accumulation of H+ and NO3 in the lichen
thalli (Boonpragob et al.. 1989). Riddell et al. (2008) exposed healthy R. menziesii thalli
to moderate (8-10 ppb) and high (10-14 ppb) HNO3 in month-long fumigations and
reported a significant decline in chlorophyll content and carbon exchange capacity
compared to thalli in control chambers. Thalli treated with HNO3 showed visual signs of
bleaching and by Day 28 were clearly damaged and dead. The damage may have
occurred through several mechanisms, including acidification of pigments and cell
membrane damage (Riddell et al.. 2008). The authors concluded that R. menziesii has an
unequivocally negative response to HNO3 concentrations common to ambient summer
conditions in the Los Angeles air basin. They believed it was very likely that HNO3
contributed to the disappearance of this sensitive lichen species from the Los Angeles air
basin, as well as other locations with arid conditions with high HNO3 deposition loads
(Riddell et al.. 2008).
Since the 2008 ISA, there have been more studies published on HNO3 effects on lichen in
the Los Angeles basin. Riddell et al. (2012) studied six lichen species with differing
3-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
morphology and physiology that were collected near the Los Angeles basin. All were
found to be sensitive to HNO3 exposures (daily peak levels near 50 ppb) in controlled
chambers. Measurements of effects included decreased chlorophyll content and
chlorophyll fluorescence, decreased photosynthesis and respiration, and increased
electrolyte leakage. The species showed differential sensitivity to the HNO3 exposures
(Riddell et al.. 2012). In the same study, lichens were not reported to be sensitive to
ozone exposure. This study adds to evidence that the main agent of decline of lichen in
the Los Angeles basin is HNO3 exposure.
In another study, Riddell etal. (2011) resampled 18 plots from a 1976-1977 study in the
Los Angeles basin. The 1976-1977 study (Sigal and Nash. 1983) had documented an air
pollution-related 50% decline of lichens described and collected in the same region in the
early 1900s (Hassc. 1913). In the 2008 resampling, Riddell et al. (2011) found
community shifts, declines in the most pollutant-sensitive lichen species, and increases in
abundance of nitrogen-tolerant lichen species compared to 1976-1977. The authors also
reported that these lichen communities have not recovered from the damaged state of the
late 1970s, and the 2008 survey data suggest that lichen communities are further
degrading. This recent observational field study further supports the evidence air
pollutants such as HNO3 may be causing declines in lichens in the Los Angeles basin.
3.5. Summary
3.5.1. Sulfur Dioxide
The current secondary standard for SO2 is a 3-hour average of 0.50 ppm, which is
designed to protect against acute foliar injury in vegetation. There has been limited
research on acute foliar injury since the 1982 PM-SOx AQCD, and there is no clear
evidence of acute foliar injury below the level of the current standard. The limited
research since 2008 adds more evidence on acute effects of SO2 on vegetation, but does
not change conclusions from the 2008 ISA on the levels producing the effects.
Effects on growth and yield of vegetation are associated with increased SO2 exposure
concentration and time of exposure. The 1982 PM-SOx AQCD concluded that more
definitive concentration-response studies were needed before useable exposure metrics
could be identified. Very few studies have been reported on the effects of SO2 on growth
of U.S. vegetation since the 1982 PM-SOx AQCD. Recent studies from eastern Europe
indicate recovery of tree growth correlated to falling SO2 concentrations since the 1980s.
El ling et al. (2009) reported that annual SO2 concentrations of 4 ppb appeared to reduce
silver fir (Abies alba) growth. There may be similar effects of SO2 emissions on trees in
June 2018
3-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
West Virginia (Thomas et al.. 2013). but more research is needed to further investigate
the mechanisms of the apparent recovery of tree growth to link this phenomenon with
declines in ambient SO2 that have occurred since the 1980s.
Limited new evidence from 2008 to the present continue to support the causal findings of
the 2008 ISA. As a whole, the body of evidence is sufficient to infer a causal
relationship between gas-phase SO2 and injury to vegetation.
3.5.2. Nitrogen Oxide, Nitrogen Dioxide, and Peroxyacetyl Nitrate
It is well known that in sufficient concentrations, NO, NO2, and PAN can have
phytotoxic effects on plants through decreasing photosynthesis and induction of visible
foliar injury (U.S. EPA. 1993). However, the 1993 Oxides of Nitrogen AQCD concluded
that concentrations of NO, NO2, and PAN in the atmosphere are rarely high enough to
have phytotoxic effects on vegetation (U.S. EPA. 1993). Since the 1993 Oxides of
Nitrogen AQCD, very little new research has been performed on these phytotoxic effects
at concentrations currently observed in the U.S.
Since the 2008 ISA (U.S. EPA. 2008a). very few studies have been published on the
direct effects of NO, NO2 and PAN on vegetation; thus, the body of evidence is
sufficient to infer a causal relationship between gas-phase NO, NO2, and PAN and
injury to vegetation.
3.5.3. Nitric Acid
The 2008 ISA reported experimental exposure of HNO3 resulted in damage to the leaf
cuticle of pine and oak seedlings, which could predispose those plants to other stressors
such as drought, pathogens, and other air pollutants (U.S. EPA. 2008a). Since the 2008
ISA, Padgett et al. (2009b) investigated dry deposition of HNO3 on the foliage, with
findings that supported the earlier research. The 2008 ISA also reported several lines of
evidence in lichen studies, including transplant and controlled exposure studies,
indicating that past and current HNO3 concentrations contributed to the decline in lichen
species in the Los Angeles basin (U.S. EPA. 2008a). Since the 2008 ISA, there have been
more exposure and field survey studies published on the effects of HNO3 on lichen in the
Los Angeles basin (Riddell et al.. 2012; Riddell etal.. 2011). These new studies continue
to support the causal findings of the 2008 ISA. As a whole, the body of evidence is
sufficient to infer a causal relationship between gas-phase HNO3 and changes to
vegetation.
June 2018
3-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
APPENDIX 4. SOIL BIOGEOCHEMISTRY
This appendix characterizes how nitrogen (N) and sulfur (S) deposition contribute to total
loading of N and S in the soils of nonagricultural terrestrial ecosystems (Appendix 4.2)
and how this loading causes soil acidification and eutrophication by altering soil chemical
pools and processes (Appendix 4.3). Additional topics are discussed, including soil
monitoring and databases (Appendix 4.4). soil biogeochemistry models (Appendix 4.5).
national-scale soil sensitivity to N and S deposition (Appendix 4.6). climate modification
of soil response to N (Appendix 4.7). and a summary (Appendix 4.8). The effects of
altered soil biogeochemistry on terrestrial biota due to soil acidification and
eutrophication are discussed in Appendix 5 and Appendix 6. respectively. Soil
biogeochemistry that has been altered by N may also be linked to aquatic
biogeochemistry, which is discussed in Appendix 7.
4.1. Introduction
The 2008 ISA for Oxides of Nitrogen and Oxides of Sulfur—Ecological Criteria
(hereafter referred to as the 2008 ISA) documented that the main effects of N and S
deposition on terrestrial soils were N enrichment and acidification. Since the 2008 ISA,
there is new evidence on how deposition contributes to total N and S loading in terrestrial
ecosystems, as well as the effects of deposition on soil chemical pools and processes.
This evidence is from addition, gradient, and time-series studies. Many of the new studies
focus on the effects of N deposition, with relatively little work focusing on S deposition.
There are improved models to evaluate ecosystem responses to deposition, most of which
are applicable at watershed scales. Some may be applied regionally. Soil N enrichment
and soil acidification occur in sensitive ecosystems across the U.S. at recent levels of
deposition. Decreasing emissions of S have led to early signs of recovery from soil
acidification in some northeastern watersheds; however, areas in the Southeast do not
show any appreciable recovery of soils. There are no signs of recovery of N enrichment
effects in soils. Critical loads (CL) determinations for soils have been made at the
ecoregion scale for NOs leaching and some soil acidification indicators. The body of
evidence is sufficient to infer a causal relationship between N and S deposition and
alteration of soil biogeochemistry in terrestrial ecosystems, which is consistent with
the conclusions of the 2008 ISA.
June 2018
4-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
4.2.
Nitrogen and Sulfur Sources to Soil
The 2008 ISA documented that atmospheric deposition is the main source of
anthropogenic N to nonagricultural and nonurban terrestrial ecosystems and headwater
streams. In 2008 it was well known that the global pool of reactive N (Nr) had increased
over the previous century, largely due to three main causes: (1) widespread cultivation of
legumes, rice, and other crops that support bacteria capable of converting diatomic
nitrogen gas (N2) to organic N through biological N fixation; (2) fossil fuel combustion
converting atmospheric N2 and fossil N to total oxidized N (NOy); and (3) the
Haber-Bosch process, which converts nonreactive N2 to Nr for N fertilizer production and
some industrial activities (Galloway et al.. 2003; Galloway and Cowling. 2002). Food
production was known to account for much of the conversion of N2 to Nr. N was shown
to be geographically redistributed through food shipment to meet human needs and often
returned to the environment via wastewater. Nr was known to accumulate in the
environment on local, regional, and global scales (Galloway et al.. 2003; Galloway and
Cowling. 2002; Galloway. 1998) in the atmosphere, soil, and water (Galloway and
Cowling. 2002). with a multitude of effects on humans and ecosystems (Townsend et al..
2003; Rabalais. 2002; van Egmond et al.. 2002; Galloway. 1998). The term the "N
cascade" was coined to refer to the sequence of transfers, transformations, and
environmental effects (Galloway et al.. 2003; Galloway and Cowling. 2002).
Since 2008, a number of estimates have been made of the relative contribution of sulfate,
oxidized N, and reduced N from atmospheric deposition. The most recent estimates are
summarized in Appendix 2. Maps showing the geographic distribution of deposition are
presented for total acidifying (N + S) deposition (Figure 2-12). total N deposition
(Figure 2-13). and total S deposition (Figure 2-31). Maps depicting how the relative
contribution of oxidized and reduced N species varies across the U.S. are presented in
Figure 2-14.
Several new studies have been published since the 2008 ISA on the source of N inputs to
ecosystems; however, no new studies on S sources to ecosystems have been identified.
Sobotaetal. (2013) quantified sources and fluxes of reactive N inputs to U.S. lands and
waterways and found human-mediated N inputs are spatially heterogeneous across the
country, ranging up to 34.6-fold the background N input across all of the 8-digit
hydrologic unit codes (HUC-8s). Across the contiguous U.S. (CONUS), synthetic N
fertilizer and atmospheric N deposition are the largest and second-largest overall
human-mediated N inputs to ecosystems, and the single largest sources in 41 and 33% of
HUC-8s, respectively (Figure 4-1 and Figure 4-2).
June 2018
4-2
DRAFT: Do Not Cite or Quote

-------
Largest human-mediated
N source
Synthetic fertilizer
Atmospheric N deposition
| Agricultural BNF
| Confined feedlot manure
| Centralized sewage
Most: Synthetic fertilizer (886 HUC-8s)
Least: Centralized sewage (32 HUC-8s)
8-digit USGS Hydrolog
USGS = U.S. Geologic Survey; HUC-8 = 8-digit hydrologic unit code; N = nitrogen; BNF = biological nitrogen fixation.
Source: map presented in Sobota et aj. (2013).
Figure 4-1 Dominant sources of nitrogen across the U.S. at 8-digit
hydrologic unit codes.
Atmospheric N deposition, circa 2000
Proportion of total N input
N = nitrogen.
Source: modified from data presented in Sobota et al. (2013).
Figure 4-2 Percentage of nitrogen input from nitrogen deposition at 8-digit
hydrologic unit codes.
June 2018
4-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
At the global scale, (Fowler etal.. 2013) estimated the relative contributions of five
sources of N that contribute to emissions and deposition, all of which are predicted to
change through the end of this century. Two new studies, both in the western U.S., have
confirmed that N from the weathering of sedimentary and metasedimentary rock can be a
source of N to some ecosystems (Montross et al.. 2013; Morford et al.. 2011). although
Sobotaetal. (2013) suggest that anthropogenic sources of N are more significant at the
landscape scale.
4.3. Soil Pools and Processes
Eutrophication and acidification are two biogeochemical processes that can occur in
response to the inputs ofN and S deposition. These processes can alter the
biogeochemistry in terrestrial ecosystems and they may occur either in sequence or
simultaneously in a given geographic area. N driven eutrophication is typically indicated
by N accumulation (e.g., increased soil N concentrations or decreased C:N ratios). These
indicators of N accumulation are directly linked to biological effects in the soil, including
changes in microbial-mediated decomposition and nitrification. N added to terrestrial
ecosystems can also be lost through leaching from the soil, typically as nitrate (NOs ). or
through emissions to the atmosphere, primarily via denitrification (Galloway et al.. 2003;
Galloway and Cowling. 2002). Denitrification is a microbial process that reduces NOs to
either unreactive N2 gas, nitrogen oxide (NO), or the potent greenhouse gas nitrous oxide
(N2O). S addition to an ecosystem typically causes S accumulation and variable amounts
of leaching from the soil, which in turn leads to acidification because demand for S as a
nutrient is low compared to human-caused soil stores of organic and inorganic S.
Soil acidification results from the accumulation of hydrogen ions (H+). This occurs
naturally through the production of carbonic acid and organic acids, as well as through
plant cation uptake (Charles and Christie. 1991; Turner etal.. 1991). The rate of soil
acidification can increase as a result of soil acidification caused by the deposition of the
inorganic acids HNO3 and H2SO4. In addition, NHx deposition contributes to soil
acidification by stimulating nitrifying bacteria that derive energy by oxidizing the NH4+
to NO3 . A byproduct of nitrification is the production of an H+ ion, but whether there is a
net effect on soil acidity depends upon the fate of the NO3 . If the NO3 is leached with a
base cation, then the H+ is left behind and the soil become more acidic. If, however, the
N03 is denitrified, then the H+ from nitrification is neutralized by OH generated by
denitrification. Likewise, if NO3 is taken up by a plant root, the root will exude an OH
in exchange.
June 2018
4-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Decreases in soil pH attributable to acid deposition have been documented throughout the
U.S. (Sullivan et al.. 2006b; Bailey et al.. 2005; Johnson et al.. 1994b; Johnson et al..
1994a). Inorganic and organic acids can be neutralized by soil weathering or base cation
exchange, in addition to denification. However, because soils vary in their capacity to
neutralize incoming acidity, the effects of acidification have been heterogeneous across
the U.S. In addition, pollutant loadings vary across the U.S. The primary chemical effects
of acidification in soils that have biological effects include the loss of important base
cation nutrients such as Ca and Mg, as well as the mobilization of aluminum (Al) cations,
several of which are toxic to many organisms. The quantities of precipitation and runoff
are important determinant of base cation leaching and acidification. (Van Breemen et al..
1984). The loss of base cations through leaching, decrease in base saturation, and
decreased in soil solution Ca:Al ratio all serve as indicators of soil acidification.
Studies published since the 2008 ISA augment our knowledge of previously identified
effects ofN addition on soils. The following sections document the empirical evidence of
N effects on multiple pools, processes, and indicators associated with the general effects
of N enrichment and eutrophication (Table 4-1). These sections summarize the empirical
effects of N and S addition on soil biogeochemistry, often based on results of from
addition or gradient studies. The publications summarized here present information on
multiple processes and indicators; therefore individual papers are often discussed in more
than one section.
4-5
DRAFT: Do Not Cite or Quote

-------
Table 4-1 Summary of key soil geochemical processes and indicators
associated with eutrophication and acidification.
Endpoint
N-Driven Nutrient
Enrichment
Acidification
Section of ISA that Discusses
Each Endpoint
PROCESS
N saturation
X
X
4.3.2
Soil N accumulation
X
X
4.3.2
NO3" leaching
X
X
4.3.2
S accumulation and adsorption

X
4.3.3
SC>42"leaching

X
4.3.3
Base cation release/depletion

X
4.3.4
Al mobilization

X
4.3.5
Nitrification
X
X
4.3.6
Denitrifi cation
X

4.3.6
Decomposition/mineralization
X
X
4.3.7 and 4.3.8
DOC leaching
X
X
4.3.9
INDICATOR
Soil [N]
X
X
4.3.2
Soil C:N ratio
X
X
4.3.6 and 4.3.7
Soil base saturation

X
4.3.4
Soil Bc:AI ratio

X
4.3.5
Fungi:bacteria ratio
X

4.3.11
Al = aluminum; Be = base cation; C = carbon; DOC = dissolved organic carbon; ISA = Integrated Science Assessment;
N = nitrogen; N03" = nitrate; S = sulfur; S042" = sulfate.
4.3.1. Nitrogen Pathways and Pools
1	The 2008 ISA documented that N is stored primarily in the soil in forest ecosystems, and
2	soil N often exceeds 85% of the total ecosystem N (Cole and Rapp. 1981; Bormann et al..
June 2018
4-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
1977). Experimental 15N addition studies showed that trees typically take up only a small
fraction of added 15N; most is retained in the soil N pool (Providoli et al.. 2005; Templer
et al.. 2005; Nadelhoffer et al.. 1999a; Tietema et al.. 1998). However, these experiments
were criticized for applying 15N directly to the soil surface, thereby precluding direct
canopy uptake of N from wet or dry deposition and limiting inference for N deposition
effects (Sievering et al.. 2000; Sievering. 1999). Forest canopies take up an average of
16% of total atmospheric N input (Lovctt. 1992). but this uptake could be considerably
higher (up to 90%) in some N limited forests with large epiphyte loads (Klopatck et al..
2006). It is unclear how much of the N from deposition retained by vegetation was used
in photosynthetic enzymes and would thus contribute to increased productivity (Bauer et
al.. 2004).
Most soil N is contained in organic matter, typically bound in organo-mineral complexes
or large-molecular-weight organic compounds (Schmidt et al.. 2011; Schimel and
Bennett. 2004) These pools of N are not directly available for uptake by plants or
microbes, and the relative immobility of these compounds means they contribute little to
the leaching loss of N into ground or surface waters. To enter the actively cycling portion
of the ecosystem N pool, recalcitrant organic N must be converted into inorganic forms
(e.g., NC>3~, NH4+) or small-molecular-weight organic compounds (e.g., amino acids,
amino sugars), typically through the activity of extracellular enzymes produced by soil
microorganisms. The size of this N pool often controls net primary productivity (NPP;
see Appendix 6). but plants compete with microorganisms for available N.
Since 2008 there is new research using 15N to trace pathways and pools in ecosystems
(Table 4-2). including a new paper confirming the previous conclusion that ecosystem N
is primarily stored in forest mineral soil (Perakis et al.. 2011). There is also new evidence
that litter is the largest sink for added N in grasslands, shrublands, and wetlands (Templer
et al.. 2012).
New studies have also been conducted tracing pathways of N in ecosystems through time.
In Swiss grasslands, N addition increased ecosystem N storage in plant biomass, both in
living plant biomass as well as litter, but did not affect soil N (Bassin et al.. 2015). In a
temperate forest ecosystem in Switzerland, application of 15N tracer in a solution to
simulate wet deposition showed increased dissolved organic N and enrichment of soil
microbes and plant roots within the first day. When plots were resampled a year later,
litter and soil pools, particularly soil organic matter, were enriched with the N tracer
(Morier et al.. 2008). These results confirm that deposited N is incorporated rapidly into
ecosystem pools and is stored primarily in litter and recalcitrant organic matter in the soil.
4-7
DRAFT: Do Not Cite or Quote

-------
Table 4-2
Pathways and pools.




Type of
Process/Indicator Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
Mineralization
Semiarid
Southern
California
Up to 35-45
forN
Two
applications
per yr of
30 kg N
N fertilizer additions resulted in faster gross N cycling
rates. Greater net release or greater net
immobilization of N was observed at different
sampling times and likely related to seasonality of C
availability.
Sirulnik et al.
(2007b)
N pools
Temperate
hardwood forest
Swiss Prealps
16

The litter layer retained approximately 19-28% of the
15N tracer after 1 day. The authors concluded that the
processes relevant for the fate of atmospherically
deposited N take place rapidly and that N recycling
within the microbe-plant-soil organic matter system
prevents further losses in the long term.
Morier et al.
(2008)
N turnover times
Semiarid
grassland
Not specified
Not specified
5 as NH4NO3
Model for N cycling was developed using
observations of15N. The temporal dynamics of 15N
fractions (labeled-N fractions) in plant and microbial
biomass are closely tied to the turnover time of these
N pools
Diikstra (2009)
N pathway
Spruce plantation
Hoglwald,
Bavaria,
Germany
45

After two to three decades of high loads of N
deposition, a new equilibrium was reached,
characterized by substantial losses of N to the
groundwater (approximately 20 kg NO3") and to the
atmosphere (16 kg N in form of N2O, NO, and N2).
Ecosystem N retention is dominated by microbial
immobilization, which was about a factor of three
higher than plant N uptake.
Kreutzer et al.
(2009)
N pathway
Simulated
northern
hardwood forest
Michigan, Great
Lakes region
7 to 12
30
Tracing 15NC>3" revealed that N accumulated in soil
organic matter by first flowing through soil
microorganisms and plants, and that the shedding of
15N-labeled leaf litter enriched soil organic matter.
Zak et al.
(2008)
June 2018
4-8
DRAFT: Do Not Cite or Quote

-------
Table 4-2 (Continued): Pathways and pools.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
N pools
Temperate forest
Oregon coast
range
2.0
None
Mineral soil accounted for 96-98% of total ecosystem
N. Soil water 15NC>3" patterns suggested a shift in
relative N losses from denitrification to NO3" leaching
as N accumulated, and simulations identified NO3"
leaching as the primary N loss pathway that
constrains maximum N accumulation.
Perakis et al.
(2011)
N pools
Soil 15N
Global
Varied
across
48 studies
Varied
across
48 studies
A meta-analysis of studies at 48 sites across four
continents shows the largest sinks for 15N tracers
among ecosystem types were organic soil in forests
(35.5%, n = 31) and foliage in tundra (12.1%, n = 3).
Litter was the largest sink in grasslands (25.5%,
n = 9), shrublands (33.8%, n = 6), and wetlands
(34.1%, n =2).
TemDler et al.
(2012)
N pools
Common garden
experiment, five
broadleaved tree
species
Denmark
13 to 19 for
broadleaf
forest
18 to 26 for
Norway
Spruce

Tree species influenced N cycling and 15N patterns
through multiple species-specific traits. The type of
mycorrhiza association, light regime, and ground
vegetation differed between ash and sycamore and
beech, lime, oak, and Norway spruce.
Callesen et al.
(2013)
Pools and pathway
N uptake
Subalpine
(seminatural)
pasture
Alp Flix, a high
plateau near Sur,
Grosons,
Switzerland
4
0, 5, 10, 25,
and 50 as
NH4NO3
Plant N pools increased by 30-40% after N addition,
while soil pools remained unaffected.
Bassin et al.
(2015)
N cycling
All
Global
Not specified
None
Synthesis of recent literature that did not include new
quantitative analysis.
Niu et al.
(2016)
June 2018
4-9
DRAFT: Do Not Cite or Quote

-------
Table 4-2 (Continued): Pathways and pools.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
Throughfall
NH4+,
NOs", and S042"
Base cations
Dry N deposition
Quercus ilex
forests ranging
from
Mediterranean
climate to a drier,
more seasonal
climate

Not specified
None
Rainfall and net NhV throughfall were negatively
correlated at all sites. Rainfall and net NC>3"-N
throughfall correlation varied between negative and
positive. Rainfall and net SO42" throughfall were
positively correlated at wet sites and negatively
correlated at the drier site.
Aauillaume et
al. (2017)
15N = Nitrogen-15; N = nitrogen; NO = nitric oxide; N20 = nitrous oxide; N02 = nitrogen dioxide; N03 = nitrate; 15N03 = Nitrogen-15-labeled nitrate; NH4N03 = ammonium nitrate;
yr = year.
June 2018
4-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
4.3.2.
Nitrogen Accumulation, Saturation, and Leaching
In the 2008 ISA, a documented indicator of terrestrial N driven eutrophication was the
accumulation of N in soil. Atmospherically deposited N can accumulate in soil as
inorganic N or by incorporating N into organic matter. The ability of atmospheric N
pollution to cause an accumulation of N in soil is indicated by a positive correlation
between the atmospheric deposition rate and total N concentration in the Oa horizon
observed at sites in New York, Vermont, New Hampshire, and Maine (Driscoll et al..
2001b). Accumulation of soil N as a result of N deposition has been either documented or
suggested to occur across large areas of the U.S. (Aber et al.. 2003) including semiarid
ecosystems (Padgett et al.. 1999). There is also evidence for soil N accumulation from
mass-balance studies of experimental N additions (Campbell et al.. 2004b'). Further
evidence that atmospheric deposition increased the availability of N in soil is provided by
the strong negative correlation between atmospheric N deposition and the decreasing C:N
ratio of the Oa soil horizon across the northeastern U.S. (Aber etal.. 2003). Soil N
accumulation is linked to increased N leaching.
New studies (Table 4-3) confirm that, across terrestrial ecosystem types, N addition
increases soil N concentrations. In semiarid shrublands in southern California, Vourlitis
and Fernandez (2012) observed that N additions increased soil N. Lu etal. (2011a)
conducted a meta-analysis of N cycle responses to N additions using data from
206 peer-reviewed studies. They observed mean increases in N leaching, soil inorganic
N, soil total N pool, as well as increases in the litter N, organic horizon N, and mineral
soil N; the only pool that decreased was microbial N (Figure 4-3).
Thresholds of N deposition associated with the onset of elevated NOs leaching have
been previously identified. Atmospheric deposition of 8 to 10 kg N/ha/yr resulted in the
onset ofNCV leaching to surface waters throughout the eastern U.S. Slightly lower N
deposition levels (5-10 kg N/ha/yr) led to NO3 leaching in the Rocky Mountains, and
this was attributable to colder temperatures, shorter growing season, slow soil
development, extensive exposed bedrock, and rapid melting of large snowpacks
(Williams and Tonnessen. 2000; Williams et al.. 1996c; Baron et al.. 1994). Lastly,
deposition loads of 17 kg N/ha/yr led to the onset of NO3 leaching in the Sierra Nevada
and San Bernardino mountains (Fenn et al.. 2008).
June 2018
4-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
N = nitrogen; Ps = photosynthesis; SIN = soil inorganic N; N-min = net N mineralization; Nit = nitrification; Den = denitrification;
SNP = soil N pool; DON = dissolved organic N.
Source: Lu et al. (2011a).
Figure 4-3 A conceptual framework for the responses of the ecosystem
nitrogen (N) cycle to nitrogen (N) addition.
The 2008 ISA documented that N saturation occurred when the input of N to the
terrestrial ecosystem exceeded the uptake capacity of the soils and biota, causing a large
fraction of the incoming N to leach from soils to surface waters (Stoddard. 1994; Aberet
al.. 1989). N saturation had been observed or suggested to occur across large areas of the
U.S. (Adams et al.. 2000; Aber et al.. 1998; Adams etal.. 1997; Peteriohn et al.. 1996;
Cook et al . 1994; Edwards and Helvev. 1991; Aber et al.. 1989).
New evidence (Table 4-3) from an N tracer study confirms that N retention varies across
ecosystem types and is highest in shrublands (89.5%) and wetlands (84.8%), followed by
forests (74.9%), and grasslands [51.8%; (Templer et al.. 2012)1. Other significant factors
affecting long-term 15N recovery (a proxy for N retention in N tracer studies) were
mycorrhizal association (cricoid > ecto > arbuscular), plant growth form, and site history
(less retention on former agricultural sites). The influence of biotic processes on N
retention is evident at smaller scales. Among nine forested sites along an urban-to-rural
landscape gradient in the Boston, MA area, throughfall inorganic N deposition increased
with proximity to the urban core, with inorganic N deposition rates positively correlated
with rates of soil inorganic N leaching across the sites (Rao et al.. 2014). Measurements
of 515N and SlsO in soil NO? leachate indicated no clear relationship between microbial
nitrification and proximity to the urban core—evidence that factors other than N
June 2018
4-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
availability influenced N processing. Overall, it is clear from recent research thatN
retention is strongly influenced by biotic factors (e.g., mycorrhizae, plant growth form)
and environmental conditions (e.g., precipitation).
Leaching of N tends to increase with increasing N addition. Where 15N tracer studies
were conducted as part of N addition experiments, the N additions decreased 15N
retention. In 15N studies with multiple N addition treatments, a negative correlation
between retention and the rate of N additions was observed (Templer et al.. 2012). In
Europe, Pise et al. (2009) documented approximately 95% of forests receiving less than
8 kg N/ha/yr still had leaching, typically less than 1 kg N/ha/yr. Additional work on
monitoring data in Sweden by Khalili et al. (2010) showed a clear sudden increase in
N03 leaching in regions where N deposition exceeded 7.5 kg/ha/yr (Appendix 4.4). In
the U.S. there are new studies modeling N leaching in eastern U.S. forests (Fakhraei et
al.. 2016; Phelan et al.. 2016).
N leaching is an indicator of ecosystem N saturation. New studies (Table 4-3) suggest the
N saturation concept (Aberetal.. 1998) may need revision in response to observations of
N cycling in temperate forests (Lovett and Goodale. 2011) and chaparral (Homvak et al..
2014). Lovett and Goodale (2011) proposed a new model of N saturation that
distinguished capacity N saturation, in which the vegetation and soil sinks for N have
been filled, from kinetic N saturation, in which the plant and soil sinks are accumulating
N, but the rate of N accumulation is slower than the N input rate. One implication of this
new model is that NOs leaching can occur even if the ecosystem N retention capacity
has not yet been saturated, as is observed at many sites (e.g., Homvak et al.. 2014;
Talhelm et al.. 2012; Lovett and Goodale. 2011). New information confirms the
applicability of the capacity N saturation concept in northwest forests (Perakis and
Sinkhorn. 2011). although the new idea of kinetic N saturation is particularly useful in
describing N cycling processes in California chaparral (Homvak et al.. 2014). At
chaparral sites, N added during the dry season is lost from the ecosystem, while N added
during the growing season is retained in plant biomass (Vourlitis and Pasquini. 2009;
Grulke etal.. 2005). In these systems, water scarcity limits plant productivity, microbial
C availability, and denitriflcation, all of which cause NO3 to be flushed from surface
soils during large precipitation events (Mcnon et al.. 2010). Large leaching losses of N in
arid ecosystems—where N demand is not indicated by biota—are evidence for kinetic
saturation.
New research highlights the importance of deposition of N from mobile sources along
roadsides (Bettez et al.. 2013). The higher deposition along roadsides in Cape Cod, MA
was associated with a two to four-fold greater rate of nitrate leaching from the soils.
Approximately 15% less N from deposition was retained 10 m from the road than sites
4-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
more distant removed from the road. N deposition contribution to N in the watershed may
be underestimated by 13-25% when roadside deposition and the associated leaching are
not included.
New research highlights the role of the microbial community in N saturation. Kopacek et
al. (2013) developed a conceptual model in which N saturation is associated with shifts in
the microbial community, namely a decrease in the fungi-to-bacteria ratio, and a
transition from N to C limitation. In N enriched systems, three mechanisms could lead to
lower amounts of bioavailable dissolved organic C (DOC) for the microbial community
and to C limitation: (1) lower plant allocation of nutrients to roots in response to
increased nutrient availability, leading to a decrease in plant exudates; (2) chemical
suppression of DOC solubility by soil acidification; and (3) enhanced bacterial
mineralization of DOC due to increased abundance of electron acceptors in the form of
N03 in anoxic soil. In support of this model, recent studies indicate that N retention in
semiarid shrublands is driven more by spatial and temporal variations in labile C
availability than exceedance ofN storage capacity (Vourlitis and Fernandez. 2015).
Furthermore, Hogberg et al. (2013) found that forest soils with low concentrations of
N03 and Al had a higher fungi:bacteria ratio compared with stands having with higher
concentrations of MVand Al (negative correlation, r = -0.857). Fungi:bacteria ratio, and
a second indicator, stem growth, explained 70% of the variation in N and Al leaching
(see broader discussion of N effects on microbial community composition in
Appendix 6).
Hogberg et al. (2013) proposed a hypothetical model to account for the effects of N
supply on plant N uptake and belowground C allocation, microbial production of
inorganic N, and N leaching (Figure 4-4). Shifts in N mineralization, nitrification, and
leaching from forests might occur in response to N loading as a result of decreasing tree
allocation of C to belowground roots, and ultimately, to ectomycorrhizal fungi and other
C limited soil microbes. When N supply increases, ectomycorrhizal fungi and other
rhizosphere microbes become progressively more C limited, and their abundance and
activity decline. Microbial assimilation of N diminishes, and N mineralization increases,
whereas the fungi:bacteria ratio decreases. Increasing NH44" levels and decreasing organic
C supply stimulate leaching, and denitrification. Increasing N loading can alter plant C
allocation, causing shifts in microbial activity and community composition that in turn
increase NOs leaching from the ecosystem (Figure 4-4).
4-14
DRAFT: Do Not Cite or Quote

-------
N supply to plants and microbes
Decreasing Increasing
I1
N leaching
C = carbon; N = nitrogen.
Thick arrows indicate an increasing rate whereas thin arrows mean decreasing rate. Nitrogen supply increases are indicated by
black arrows.
Source: Hoaberg et al. (2013).
Figure 4-4 A hypothetical model to account for the effects of nitrogen supply
on plant nitrogen uptake and belowground carbon allocation,
microbial production of inorganic nitrogen, and nitrogen leaching.
June 2018
4-15
DRAFT: Do Not Cite or Quote

-------
Table 4-3 Nitrogen accumulation, saturation, and leaching.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
N leaching
Forest	248 sites (plot and
catchment scale)
from 15 countries in
Europe
Varied
Not reported Gradient: The most consistent indicators of N
leaching were throughfall N deposition, organic
horizon C:N ratio, and mean annual temperature.
Sites receiving low levels of N deposition
(8 kg N/ha/yr) showed very low output fluxes of N.
In general, the models successfully predicted N
leaching (mean of ± 5 kg N/ha/yr between
observed and predicted) from forests at early to
intermediate stages of N saturation but not from N
saturated sites.
Pise et al.
(2009)
N accumulation
N leaching
Mesic desert
Spanish Spring
Valley, NV
Not specified None
Field Experiment: Concluded that vadose soil
resources (water and organic C) are rare. Unused
NO3" from low biological demand is transported
and accumulated in the deeper vadose zone with
occasional deep leaching events.
Menon et al.
(2010)
N saturation
Oak forest Southeastern New 9
York State
100 NH4NO3
(1996-1999)
50 NH4NO3
(2000-2006)
Conceptual Model: New N saturation model
based on an N addition study of an oak forest in
southeastern New York State.
Lovett and
Goodale (2011)
N accumulation
meta-analysis Varied
Varied
Meta-Analysis: N cycle responses to N additions
using data from 206 peer-reviewed studies and
observed a mean increase in N soil pools except
microbial biomass N (Fiaure 4-12).
Lu et al. (2011a)
June 2018
4-16
DRAFT: Do Not Cite or Quote

-------
Table 4-3 (Continued): Nitrogen accumulation, saturation, and leaching.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
N saturation
N accumulation
N mineralization
nitrification
Douglas-fir
forest
Oregon coast range 2.0
None	Gradient: This is a test of the N saturation theory Perakis and
along a natural N gradient. Higher N content of Sinkhorn (2011)
surface (0-10 cm) soil was linearly related to
higher net N mineralization. Lower pH was related
to lower nitrification. The ratio of netgross N
mineralization to higher nitrification increased
along the gradient, indicating progressive
saturation of microbial N demand.
N leaching	Douglas-fir Oregon coast range 2.0
forest
None	Gradient: Hydrologic N losses were dominated by Perakis and
dissolved organic N at low N sites, with increased Sinkhorn (2011)
nitrate loss causing a shift to dominance by nitrate
at high N sites, particularly where net nitrification
exceeded plant N demands.
N leaching 48 sites across Global	Varied across Varied across Meta-Analysis: The greatest recoveries of	Templer et al.
four continents	48 studies 48 studies ecosystem 15N tracer occurred in shrublands (2012)
Grassland	(mean, 89.5%) and wetlands (84.8%) followed by
forest wetland	forests (74.9%) and grasslands (51.8%).
shrubland
Soil N, C:N
Chaparral and
coastal sage
scrub (CSS)
Southern California 6-8.1
56 to 58	Addition: In this 6-yr field experiment, chaparral Vourlitis and
and CSS vegetation communities were found to Fernandez
have the capacity to immobilize 6.2 and 11.9 g (2012)
N/m2/yr, respectively. Soil extractable N increased
significantly after 7-10 g/m2 of cumulative N
exposure, resulting in a simultaneous increase in
the N concentration and a decline in the C:N ratio
of shrub tissue. Similar results were observed for
the surface litter pool and litter production but at a
higher cumulative N exposure.
June 2018
4-17
DRAFT: Do Not Cite or Quote

-------
Table 4-3 (Continued): Nitrogen accumulation, saturation, and leaching.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
NO3 leaching
Forest
Cape Cod, MA	Annual	None	Roadside Gradient: 25 to 30% higher N
throughfall	deposition to forests 10 m away from roads on
deposition was	Cape Cod, compared to hundreds of meters away
8.7	(±0.4) and	from the road; the higher deposition was
6.8	(±0.5) TDN	associated with a two to four-fold greater rate of
at sites 10 and nitrate leaching from the soils. 73% of the
150 m away	deposition was retained in the forest away from
from the road	roads, compared to 58% retention at 10 from the
road. Results scaled to the entire watershed
indicate an underestimate of the amount of N
deposition to the watershed by 13-25% by not
including roadside deposition and leaching.
Bettez et al.
(2013)
Soil [NO3-]
Soil [Al]
Fungi:bacteria ratio
19 Picea abies South Sweden	Throughfall N Ammonium Addition: Microbial community composition in the
(L.) Karst.	includes wet NC>3"at20 organic horizon and soil solution chemistry below
Stands	and dry inputs	the rooting zone were highly correlated. Stands
and ranged from	with low concentrations of NO3" and Al had higher
2.7 to 19	fungi:bacteria ratio compared with stands with
higher concentrations of NO3" and Al. Microbial
community composition in the soil was more
closely related to the soil solution than to the soil
chemistry. The study found a significant negative
correlation between the fungi:bacteria ratio in the
soil and NO3" and Al in soil solution (r = -0.533
and -0.857, respectively).
Hoqberq et al.
(2013)
N saturation
None
(theoretical)
None (theoretical) Not specified None
Conceptual Model: N addition alleviates N
limitation, and together with SO42" deposition,
causes soil acidification and increases availability
of electron acceptors for soil microbial processes.
Increasing N and SO42" decreases fungal
biomass, increases bacterial DOC mineralization,
and decreases DOC leaching.
Kopacek et al.
(2013)
Decomposition
Deciduous
forests
Catskills Mountains
of southeastern New
York
9.0
50 (NH4NO3) Addition: NO3" leaching increased markedly in Lovett et al.
response to the N addition in all species,	(2013)
indicating that the addition rate of N exceeded the
N retention capacity of vegetation and soils in
these plots.
June 2018
4-18
DRAFT: Do Not Cite or Quote

-------
Table 4-3 (Continued): Nitrogen accumulation, saturation, and leaching.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
NO3" leaching
Alpine
Rocky Mountain
National Park (Niwot
Ridge)
Not specified
Variable
Review: Nitrate leaching increases above 10 kg
N/ha/yr deposition.
Bowman et al.
(2014)
Rate of soil N cycling
Alpine
Rocky Mountain
National Park (Niwot
Ridge)
Not specified
Variable
Review: the rate of soil N cycling increases above
15 kg/ha/yr deposition.
Bowman et al.
(2014)
Soil acidification and
soluble Al
Alpine
Rocky Mountain
National Park (Niwot
Ridge)
Not specified
Variable
Review: soil acidification and soluble aluminum
increases above 25 kg/ha/yr deposition.
Bowman et al.
(2014)
N export	Forest	Ontario, Canada Not reported Not applicable Isotope: Rain on snow, as it passes through the Casson et al.
ecosystem, has a higher concentration of NO3" (2014b)
(throughfall/snowmelt average = 498 pg/L)
compared with baseflow (average = 7.3 |jg/L;
average = 41 pg/L) and as a result, throughfall
and snowmelt contribute the majority of NO3"
export (average = 62%) during rain on snow
events.
Gradient: The source of N leaching from five of Rao et al. (2014)
nine sites was almost entirely from nitrification,
indicating that the NCVin leachate came from
biological processes rather than directly passing
through the soil. A significant proportion
(17-100%) of NC>3"leached from the other four
sites came directly from the atmosphere.
6.6 to 6.7 N Not applicable Isotope: There was no significant biological	Templer et al.
8.1 to 8.5 SO42"	production of NO3" via nitrification in the canopy. (2015b)
NO3" concentrations in streams were low and had
natural 1sO abundances consistent with microbial
production, demonstrating that atmospheric N is
being biologically transformed while moving
through these watersheds and that these forested
watersheds are unlikely to be N saturated.
NO3 leaching	Urban gradient	Urban sites 12.3
and nonurban
5.7
N saturation/NhU , Forest	White Mountains
NO3", total inorganic	National Forest,
N, and SC>42"-S	Crawford Notch, NH;
Lye Brook, VT in
southern Green
Mountain National
Forest
June 2018
4-19
DRAFT: Do Not Cite or Quote

-------
Table 4-3 (Continued): Nitrogen accumulation, saturation, and leaching.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
N accumulation
Semiarid
chaparral and
CSS
Santa Margarita
Ecological Reserve
and the Sky Oaks
Field Station
6 to 8
50
Addition: N enrichment significantly increased N
accumulation but not microbial respiration.
Vourlitis and
Fernandez
(2015)
N retention
Beech and
Norway spruce
forests
Soiling, Germany
Direct total None	Field Time Series: Two study plots observed that Meesenburq et
deposition not	N retention decreased from 40 to 0-20 kg/ha/yr al. (2016)
specified	(from 1970s to current), indicating increasing N
saturation.
NO3 leaching, and N
availability
Eastern U.S.
forests
Hubbard Brook
Experimental Forest,
NH, and Bear Brook
Watershed, ME
Four modeling
scenarios
ranging from 7.4
to 103.9 keq/ha
N and 7.8 to
149.1 keq/ha S
from 1850-2100
with 0-95%
decrease from
2010 to 2100
28.8 kg/hr/yr Model: Modeled scenarios at Hubbard Brook, NH
S and	received twice the N and S at Bear Brook, ME
25.2 kg/ha/yr and resulted in the largest changes in soil base
(NH^SCM saturation, ANC, NO3" leaching and N availability.
Modeled recovery of soil chemistry and
understory plant communities at both forests only
occurred when N and S deposition were modeled
at preindustrial levels.
Phelan et al.
(2016)
NO3 leaching
Mixed
beech-pine
forest
Neuglobsow
Integrated
Monitoring (IM) site,
Germany
Total N wet
deposition:
13.26 ±2.01 kg
N/ha/yr
(1998-2013)
None	Field Observation: Seepage water (120 cm) is
estimated to contain at 2.38 kg N/ha/yr (96% as
NO3-).
Schulte-Bispinq
and Beese
(2016)
N immobilization
Restored
tallgrass prairie
Prairie invasion and
Climate Experiment
(PRICLE) Loveland,
CO
-1/3 of N added
5 g N/m2/yr
urea.
"Seasonal
maximum
5-day
cumulative
rainfall"
increased by
33% (2012)
and 9%
(2013)
N Addition: Nitrogen addition during plant growth
accelerated subsequent mass loss of
Schizachyriumscoparium litter and litter produced
with N addition had a 65% greater N loss than
ambient N litter in Solidago Canadensis plots,
indicating N addition accelerates N cycling.
Schuster (2016)
June 2018
4-20
DRAFT: Do Not Cite or Quote

-------
Table 4-3 (Continued): Nitrogen accumulation, saturation, and leaching.
Process/Indicator
Type of
Ecosystem
Region
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Effect of Deposition
Reference
HERO ID
N soil retention
Forest
Great Smoky
Mountains National
Park
5.1 kg N/ha/yr None	Model: PnET-BGC was used to model 30 stream
(36.5 mmolc	watersheds characterized by decreased SO42"
/m2/yr),	and NO3" deposition during 1981-2014 (81 and
respectively.	53%) and predict stream recovery. Spruce-fir
forests at higher elevations have limited N
retention and exhibit N saturation due to elevated
N deposition.
Fakhraei et al.
(2016)
Al = aluminum; C = carbon; cm = centimeter; CSS = coastal sage scrub; DOC = dissolved organic carbon; g = gram; ha = hectare; kg = kilogram; L = liter; m = meter; N = nitrogen;
NH4+ = ammonium; N03 = nitrate; P = phosphorus; NH4NO3 = ammonium nitrate; r = correlation coefficient;; S042 = sulfate; yr = year.
June 2018
4-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
4.3.3.
Sulfate Accumulation, Adsorption, and Leaching
The 2008 ISA showed that SOx deposition may be assimilated by vegetation or microbes,
accumulate in the soil, or act as a mobile ion and leach out of the soil to aquatic
ecosystems. Plant demand for S is typically low, particularly in comparison to the large
pools of S stored in the soil in recalcitrant organic compounds; consequently almost all S
deposited in terrestrial ecosystems enters soil rather than plant pools. At many locations
in the U.S. that receive high levels of S deposition, notably the glaciated Northeast and
upper Midwest, much of the deposited S leaches through soils into streams and lakes. The
physical process of charge balance pairs sulfate (SOr ) leaching with leaching of
countercharged cations from the soil matrix, and this process contributes to acidification
of soil, soil water, and surface water. As the base cations become depleted in the soil
matrix, charge balance in the soil is maintained by an increase in acidic cations (H+ and
inorganic Al), sometimes resulting in toxic conditions for plant roots and aquatic
organisms (Charles and Christie. 1991; Turner et al.. 1991).
In the 2008 ISA, regional trends of SOr soil accumulation and leaching were identified
in the U.S. In the Southeast, accumulation of atmospherically deposited S in soil resulted
from S042 adsorption to soil particles as well as incorporation of S into soil organic
matter through biological assimilation. Accumulated S is slowly released from soil pools
into drainage water, and this process can temporarily delay ecosystem recovery in
response to decreases in S deposition (Sullivan et al.. 2004; Elwood et al.. 1991; Turner
et al.. 1991). In the Northeast, there was a demonstrated accumulation of S in soil
(Driscoll et al.. 2001a). Two new studies on SO42 accumulation, leaching, and
adsorption are described below (Table 4-4).
The net loss of S from soils is occurring in a number of northeastern watersheds in
response to decreased levels of atmospheric S deposition. In a new study evaluating
watersheds in the Northeast, Mitchell and Likens (2011) calculated that annual
discrepancies in the watershed S budgets (SO42 flux in drainage waters minus total
atmospheric S deposition) have become more negative, indicating the increasing
importance of the release of S from internal ecosystem sources. The release of S from
forest soils is controlled (57%) by water flux and soil moisture(Mitchell and Likens.
2011).
In the southeastern U.S., Rice et al. (2014) calculated SO42 mass balances for
27 forested, unglaciated watersheds from Pennsylvania to Georgia by using total
atmospheric deposition (wet plus dry) as input. Unlike their counterparts in the
northeastern U.S. and southern Canada, many of these watersheds still retain SO42 . Rice
June 2018
4-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
et al. (2014) predicted that many of the watersheds in the study will begin releasing SO42
over the next two decades. The specific years when the watersheds cross over from
retaining to releasing SO42 correspond to a general geographical pattern from north to
south of later net watershed release. For instance, the three watersheds in West Virginia
have crossover years that ranged from 2006 to 2011, and crossover years for five of the
watersheds in Virginia range from 2012 to 2021. The runoff ratio, computed as the ratio
of annual mean discharge to annual mean precipitation, was the single best
watershed-scale predictor of the crossover year (r2 = 0.72). Watersheds with higher
runoff ratios tend to convert sooner from net retention to net release of SO42 (Rice et al..
2014). More recently, Fakhraei etal. (2016) used PnET-BGC to model 30 stream
watersheds during 1981-2014 in the Great Smoky Mountains National Park. Hindcast
modeling (beginning ca. 1850) to 2014 showed that the soil SO42 pools increased from
20.2 g/m2 (preindustrial) to 145.6 g/m2 (ambient median). Soils with high SO42
adsorption capacity also had a faster rate of base cation depletion.
4-23
DRAFT: Do Not Cite or Quote

-------
Table 4-4 Sulfate adsorption, accumulation and leaching.
Process/Indicator Type of Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Sulfate adsorption Upland forests,
and cation	low-lying areas and
leaching	wetlands
Athabasca oil sands Not specified
region (AOSR) in
Alberta, Canada
None
Field Comparison: SO42"
adsorption capacity was relatively
low (50 to 500 mg SC>42"/kg) in both
watersheds as compared to other
acid-sensitive soils in eastern North
America.
Jung et al. (2011)
S budget
Varied
15 sites
southeastern
Canada and
northeastern U.S.
-0.1 to 14
Long-Term Deposition: The net
annual fluxes of SO42" showed a
strong relationship with hydrology;
the sensitivity of S budgets is likely
greatest in watersheds with the
greatest wetland area, which are
particularly sensitive to drying and
wetting cycles.
Mitchell et al. (2011)
Sulfate
accumulation and
leaching
Forest
HBEF in the White
Mountains of New
Hampshire
-7 to 20
None
Long-term Deposition: S released
from internal sources is increasing
overtime. Watershed wetness, as a
function of log-io annual water flux
explained 57% (n = 157) of the
annual variation for four
watersheds. The biogeochemical
control of annual SO42" export in
streamwater draining from forested
watersheds has shifted from control
by atmospheric S deposition to soil
moisture driven by climate.
Mitchell and Likens
(2011)
June 2018
4-24
DRAFT: Do Not Cite or Quote

-------
Table 4-4 (Continued): Sulfate adsorption, accumulation and leaching.
Process/Indicator Type of Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
SO42 leaching
Watersheds ranging
from Pennsylvania to
Georgia
27 forested,
unglaciated
watersheds
Not specified
None
Long-term Deposition: Calculated
SC>42"mass balances for
27 watersheds showed that over
the next two decades, many of the
study's watersheds will begin
releasing SO42". The single most
important variable that explained
the crossover year was the runoff
ratio, defined as the ratio of annual
mean stream discharge to
precipitation.
Rice etal. (2014)
Soil SO42"
Forest
Great Smoky
Mountains National
Park
3.1 kg S/ha/yr
(19.3 mmolc/m2/yr)
None Model: PnET-BGC used to model
30 stream watersheds during
1981-2014 when SO42" and NOs"
deposition decreased (81 and 53%,
resp.). Hindcast modeling
(beginning ca. 1850) increased soil
pools of SO42" from 20.2 g/m2
(preindustrial) to 145.6 g/m2
(current median).
Fakhraei et al. (2016)
Base saturation
Forest
Great Smoky	5.1 kg N/ha/yr
Mountains National (36.5 mmolc/
Park, U.S.	m2/yr)
None Model: PnET-BGC used to model
30 stream watersheds during
1981-2014 when S042" and NOs"
deposition decreased (81 and 53%,
resp.). High capacity SO42"
adsorbing soils depleted base
saturation faster.
Fakhraei et al. (2016)
Soil S042"
Boreal forest
Sweden
Not specified	None Time Series: In an analysis of 10 yr Ledesma et al. (2016)
of data, riparian zone soil SO42"
was observed to decrease from
2003-2012.
June 2018
4-25
DRAFT: Do Not Cite or Quote

-------
Table 4-4 (Continued): Sulfate adsorption, accumulation and leaching.
Process/Indicator Type of Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Soil SO42"
Maritime pine forest
plantation
Northwest Spain Not specified
None Field Experiment: The soil parent
material (slate, biotitic schist, mica
schist, and granite) did not
significantly affect soil solution
SO 42"(p = 0.39).
Eimil-Fraaa et al.
(2016)
AOSR = Athabasca oil sands region; S042 = sulfate; kg = kilogram; HBEF = Hubbard Brook Experimental Forest; S = sulfur; yr = year.
June 2018
4-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
4.3.4.
Base Cation Leaching and Exchange
In the 2008 ISA, it was known that acidifying deposition changes concentrations of
exchangeable base cations in soil by accelerating natural rates of base cation leaching
until stores become depleted (Lawrence et al.. 1999; Cronan et al.. 1978). Base cations
include the essential plant nutrients (e.g., Ca, Mg, and K), and the loss of exchangeable
base cations from the soil may have adverse effects on flora. When SO42 and NO3
leaching occur in equal magnitude to base cation leaching, the drainage water is not
acidified. However, in the process of neutralizing the acidity of drainage water, base
cation release from soil eventually can cause decrease of the base saturation of the soil.
Soil base saturation expresses the concentration of exchangeable bases (Ca, Mg,
potassium [K], sodium [Na]) as a percentage of the total cation exchange capacity (which
includes exchangeable H+ and inorganic Al).
Under conditions of low soil base saturation (approximately <20%) and elevated
concentrations of strong acid anions, Al is mobilized from soil to drainage water (Cronan
and Schofield. 1990; Reuss. 1983). with potentially harmful consequences for sensitive
terrestrial plants and aquatic organisms (Appendix 7) throughout the food web
(Appendix 4.3.6).
Leaching of base cations by acidifying deposition has been documented in sensitive
regions in the U.S., including the Adirondack Mountains, New England, the Catskill
Mountains, and northwestern Pennsylvania (U.S. EPA. 2008a). Base cation loss increases
the sensitivity of the watershed to further acidifying deposition. Watersheds that were
capable of fully neutralizing a particular level of acidifying deposition in the past may no
longer be capable of fully neutralizing that level today or in the future due to the
cumulative effect of acidifying deposition on soil base saturation.
Base saturation values less than 10% predominate in the soil B-horizon in the areas in the
U.S. where soil and surface water acidification from acidifying deposition have been
most pronounced, including conifer and hardwood forests in the Adirondack Mountains
(Sullivan et al.. 2006b). red spruce forests throughout the Northeast (David and
Lawrence. 1996). hardwood forests in the Allegheny Plateau (Bailey et al.. 2004). and
conifer and hardwood forests in the southern Appalachian Mountains (Sullivan et al..
2003). In a study of sugar maple decline throughout the Northeast, Bailey et al. (2004)
found threshold relationships between base cation availability in the upper B soil horizon
and sugar maple mortality at Ca saturation less than 2%, and Mg saturation less than
0.5% (Bailey et al.. 2004). The authors concluded that base saturation varied as a function
of topography, geologic parent material, and acidifying deposition.
June 2018
4-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
New publications further support findings from the 2008 ISA that N and S deposition
cause base cation depletion from soils (Table 4-5). In the Rocky Mountains, Lieb et al.
(2011) observed that soil acid buffering capacity decreased as N inputs increased (40%
decrease at highest N input). An acidification threshold was calculated for significant loss
of soil acid buffering capacity of around 28 kg N/ha/yr. In another U.S. study, long-term
trends in base cation depletion at Bear Brook watershed, ME, showed N and S addition
over a 17-year time period (while ambient S deposition was simultaneously decreasing)
resulted in little evidence of continued soil exchangeable base cation depletion or
recovery [expected because of decreasing S deposition (SanClements et al.. 2010)1. A
study of the forest understory herb community in West Virginia found that N addition
lowered plant-available Ca and, to a lesser degree, Mg, but not K, illustrating how
biogeochemical cycling of forest ecosystems is altered (Gilliam et al.. 2016a).
A meta-analysis of 107 studies found N addition alters the availability of base cations in
terrestrial and aquatic ecosystems (Lucas etal.. 2011): although short-term N and S
deposition cause base cation depletion, long-term trends across all studies are unclear and
may be affected by confounding disturbances. Evaluating the strength of these results is
difficult because they are based on averages from various biome types and there are few
long-term studies.
Field studies in forests in Europe confirm that N deposition and N addition lower soil pH
and decreases base cations (Chen et al.. 2015; Ferretti et al.. 2014). As acidifying
deposition decreased, base cation concentration in the soil increased (Berger et al.. 2016).
Two studies from grasslands in Asia report mixed results, with N addition causing Be
levels to decrease in one study (Chen et al.. 2015) and increase in another (Tian et al..
2016b). A study of European grasslands found that base cation depletion increased with
N addition over a 10-year period, leading to a loss of 23 to 35% of total available bases
(Ca2+, Mg2+, K+, and Na+) from the soil and acidifying it by 0.2 to 0.4 pH units (Horswill
et al.. 2008).
Base cation weathering rates are uncertain, but substantial advancements have been made
in this field since the 2008 ISA (Appendix 4.5.1.1). New model estimates have been
published to two forested areas in Canada (Williston et al.. 2016; Watmough et al.. 2014).
Major sources of sources of Be to ecosystems are either from atmospheric deposition or
weathering from soils. Two additional studies from Spain considered the sources of Be in
ecosystems from deposition (Aguillaume et al.. 2017) or as compared with sources from
different types of soil parent materiaKEimil-Fraga et al.. 2016).
Additional literature that evaluates how base cation depletion from acidification may
recover in response to the addition of base cations to the soil is also noted here; however,
4-28
DRAFT: Do Not Cite or Quote

-------
1	this literature does not describe the effects of N and S deposition, but rather a method for
2	recovering ecosystems to a more natural state. This literature includes several
3	publications from a 15-year Ca addition study at Hubbard Brook Experimental Forest,
4	NH (Shao et al.. 2016; Johnson et al.. 2014; Green et al.. 2013; Nezat et al.. 2010).
June 2018
4-29
DRAFT: Do Not Cite or Quote

-------
Table 4-5
Base cation leaching and exchange.


Process/
Indicator
Type of Deposition Addition
Ecosystem Region kg/ha/yr kg/ha/yr
Effect of Deposition
Reference
HERO ID
Addition: Treatments caused grassland soils to lose 23 to 35% of Horswill et
their total available bases (Ca2+, Mg2+, K+, and Na+), and they al. (2008)
became acidified by 0.2 to 0.4 pH units. Al, Fe, and Mn were
mobilized and taken up by limestone grassland forbs and were
translocated down the acid grassland soil. Mineral N availability
increased in both grasslands and many species showed foliar N
enrichment. N deposition depletes base cations from grassland
soils.
Base cation Grasslands Peak District Not specified Plots treated for
depletion	National Park,	8 to 10 yr with 0,
Soil pH	England	35, or 140 N as
Soil [Al], [Fe],
[Mg]
NH4NO3
Base cation
depletion
Hardwood
forest
Eastern U.S.,
Bear Book
watershed, ME
28.8 S and 25.2 Addition: Compared treated and untreated watersheds after N
N as (NH4)2SC>4 and S manipulation over a 17-yrtime period. Found little evidence
of continued soil exchangeable base cation concentration
depletion or recovery, possibly because a 1998 ice storm
increased litterfall and accelerated mineralization, obscuring
temporal trends in soil chemistry.
SanCleme
nts et al.
(2010)
Base cation
depletion
Critical load
Forest stands
with mature
white ash
Niwot Ridge in
southern
Rocky
Mountains
20, 40, 60 N Addition: Soil acid buffering capacity decreased with increasing N Lieb et al.
inputs (40% decrease at highest input), and was associated with a (2011)
decrease in pH, loss of extractable Mg2+ and increases in Mn and
Al3+. The threshold at which acidification occurred was around
28 kg N//ha/yr.
Base cation Boreal forest, 107 sites
depletion	temperate globally
forest, tropical
forest, and
grassland
Not specified
Median N
addition across
the studies was
38, and 71% of
the studies
added 70 or
less
Meta-analysis: Evaluation of 107 independent studies to
determine whether N fertilization alters the availability of base
cations (Be) in terrestrial and stream ecosystems. Results suggest
N fertilization may accelerate Be loss from terrestrial ecosystems
over time periods less than 5 yr.
Lucas et al.
(2011)
June 2018
4-30
DRAFT: Do Not Cite or Quote

-------
Table 4-5 (Continued): Base cation leaching and exchange.
Process/
Indicator
Type of
Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Base cation Jack pine	Athabasca oil Not specified 30 N, 30 S Addition: No evidence of N saturation in the studied forest
leaching (Pinus	sands region forAOSR (2006 through ecosystem after 4 yr of N and S additions. No long-term increase
banksiana) and	(AOSR), 2009) of inorganic N concentrations in the soil; leaching of N beyond the
aspen	Alberta, main rooting zone in the soil profile was minimal, and tree growth
(Populus	Canada was increased by N addition, all indications of N limitation in the
tremuloides) in	studied forest stand. However, exchangeable Ca2+ and Mg2+
upland forests	concentrations in the surface mineral soil layer were reduced by N
and black	and S additions because of increased cation leaching associated
spruce (Picea	with increased SO42" leaching caused by S addition and
mariana) in	increased nutrient uptake associated with increased tree growth
low-lying areas	resulting from N addition,
and wetlands.
Jung and
Chang
(2012)
Hydrologic Forest	Hubbard Brook
flowpath	Experimental
Forest, NH
watershed
Addition: The flow path of the added Ca was followed through (Nezatet
time. The deepest flowpaths to the streams were penetrated by al. (2010)
3-9 yr after application. It was estimated that only -360 kg out of
19 metric tons of Ca applied as wollastonite had been exported
from the watershed in stream flow 9 yr after its application and it
would take 1,000 yr for all of the added Ca to be transported from
the watershed.
Not specified
41 metric tons
of the mineral
wollastonite
(CaSiC>3) was
applied to an
11.8 hectare
Evapotran- Forest	Hubbard Brook Not specified 1,028 kg Ca/ha
spiration	Experimental	(as wallastonite)
Forest, NH	in 1999
Field Observation: 25, 18, and 19%, evapotranspiration,	Green et
increased in Years 1-3 after addition and then returned to	al. (2013)
pretreatment levels. Watershed soil retained Ca from the
wollastonite, indicating a watershed-scale fertilization effect on
transpiration. That response is unique in being a measured
manipulation of watershed runoff attributable to fertilization.
June 2018
4-31
DRAFT: Do Not Cite or Quote

-------
Table 4-5 (Continued): Base cation leaching and exchange.
Process/
Indicator
Type of
Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Total Ca,
exchangeable
Ca, cation
exchange
capacity, base
saturation
Forest	Hubbard Brook Not specified 1,028 kg Ca/ha Field observation: Compared to conditions before the 1999 Ca Johnson et
Experimental	(as wallastonite) addition, total Ca increased 596% in the Oie-horizon in 2000, 81% al. (2014)
Forest, NH	in 1999	in the Oa-horizon in 2002, and 146% in the 0-10 cm mineral soil
in 2006.
Oie-horizon exchangeable Ca tripled in 2000 and remained
significantly higher through 2010. Significant increases occurred in
2002 in Oa; modest increases measured in upper mineral soil in
2006 and 2010. Total Ca pool in O-horizon and mineral soils was
250 kg/ha greater than pretreatment. In total, 92.4% of Ca added
was estimated to dissolve and enter the ecosystem as labile Ca.
Cation exchange capacity CEC increased by 106% in the
Oie-horizon in 2010 but decreased 52 and 32% in the Oa and
mineral soil. Oie, Oa, and mineral soil base saturation rose 69, 84,
and 58%, resp. by 2010.
PH
Base cations
Forest soils
Italy
4.5 to 28.8 Not applicable
throughfall N
(NOs" +
NH4+)
Gradient: Exchangeable base cations and pH decreased with
Ferretti et
increasing N deposition, and foliar nutrient N ratios (especially N:P al. (2014)
and N:K) increased. Comparison between bulk openfield and
throughfall data suggested possible canopy uptake of N, levelling
out for bulk deposition >4-6 kg/ha/yr.
Be weathering Boreal plains NE Alberta, S = 118,	Not applicable Modeling: base cation weathering rate estimated for 63 sites
rate Canada NhV = 93,	using PROFILE was 17 mmolc m2/yr, however acidification was
NO3" = 49 in	not expected because base cations from fugitive dust sources
units of	were a comparatively high 250 mmolc m2/yr, offsetting much of
mmolc m2/yr	the acidifying input of N and S deposition.
Watmouqh
et al.
(2014)
PH
Soil N
Base cations
Fungi:bacteria
ratio
Belowground
biomass
Microbial
community
Structure
Semiarid	Mongolia	Not specified 0,17.5,52.5, Addition: Soil pH decreased across the N addition gradient by
grassland	105.0, 175, and 0.3-1.8 units in 2010 and by 0.1-1.7 units in 2011. Decreased
280 NH4NO3 concentrations of mineral cations Ca2+, Mg2+, and Na+ were
fertilizer	observed. The observed increases in above- and belowground
biomass and changes in plant community structure were mainly
(57-69%) attributed to the increase in soil N availability and
changes in soil base cations. N addition increased the
fungi:bacteria ratio by 5-18% in 2010 and by 2-10% in 2011.
Chen et al.
(2015)
June 2018
4-32
DRAFT: Do Not Cite or Quote

-------
Table 4-5 (Continued): Base cation leaching and exchange.
Process/	Type of	Deposition Addition
Indicator Ecosystem Region	kg/ha/yr	kg/ha/yr
Base	Adirondack SC>42"= None
saturation	Mtns. region, 290.3-365.7
NY	eq/ha/yr;
N03" = 172.5
-233.5
eq/ha/yr
Reference
Effect of Deposition	HERO ID
Model: PnET-BGC was used to evaluate biophysical factors that Zhou et al.
affect CLs of acidity. Model simulations included a range of future (2015c)
scenarios of decreases in atmospheric nitrate, ammonium, and
sulfate deposition from the present to 2,200; historical forest
harvesting; supply of naturally occurring organic acids; and
variations in lake hydraulic residence time.
Assuming the current soil base saturation of 6.4%, a decrease in
SO42" deposition from 0 to 100% resulted in a soil percentage BS
range of 6.2 to 15.3%, respectively, after 200 yr. (Preindustrial soil
percentage BS -22%).
Be	Predominantly ViennaWoods, Not specified None	Time Series: In 1984 and 2012, soil samples were taken from Beraer et
beech	Austria	locally	20 cm downhill and 3 m away from the base of a beech tree stem. al. (2016)
Exchangeable Ca2+, Mg2+, and pH increased in 0-5 cm soil from
1984 to 2012. Recovery appeared delayed in deeper soils. Foliar
base cations Ca, Mg, and K decreased. (Foliar K declined the
most at 48%).
Be
Maritime pine
Northwest
Not specified
None
Field observation: The soil parent material (slate, biotitic schist,
Eimil-Fraqa

forest
Spain


mica schist, and granite) significantly affected base cation
et al.





concentrations.
(2016)
Foliar base
Eastern U.S.
Fernow
-10 kg
35 kg N/ha/yr
N Addition: Foliar measurements were used as a proxy for soil
Gilliam et
cation
temperate,
Experimental
N/ha/yr (wet)
(NH4)2S04
micronutrient availability under N addition. Excess N lowered
al. (2016a)

hardwood
Forest, WV


plant-available Ca and, to a lesser degree, Mg, but not K. N


forest



addition significantly affected Ca:AI ratios in Viola sp and Rubrus


herbaceous



sp.


layer (Viola






rotundifolia,






Rubus






allegheniensis)





June 2018
4-33
DRAFT: Do Not Cite or Quote

-------
Table 4-5 (Continued): Base cation leaching and exchange.
Process/
Indicator
Type of
Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Soil solution Be Forest
(2016)
Hubbard Brook Not specified 1,028 kg Ca/ha Field Observation: Significant increase in soil Ca concentrations. Shao et al.
Experimental	(as wallastonite) Increases in pH and ANC and decreases in inorganic Al in the Oa,
Forest, NH	in 1999	Bh, and Bs soil horizons. Ca:AI ratios ranging from 1 to 5.3 in the
Oa-, Bh-, and Bs-horizons before treatment increased significantly
after Ca addition, with the greatest increase occurring in the
Oa-horizon. Average 2000-2011 Ca fluxes in soil solution across
the study area's three subwatersheds ranged from
26.4 + 3.2 mmol/m2/yrto 29.9 + 4.2 mmol/m2/yr. The weighted
average percentage increase in Ca fluxes for the three
subwatersheds from 1999-2011 were 139, 91, and 97%forOa-,
Bh-, and Bs-horizons, respectively.
Soil Be
Temperate
grassland
Mongolia,
China
1.6 mg
N/m2/yr (not
specified if
wet or total)
1, 2, 4, 8, 16,
32, 64 g N/m2/yr
(urea)
N Addition: Soil exchangeable Mn2+, Fe3+, and Al3+
concentrations increased linearly with N addition.
Tian et al.
(2016b)
Be weathering
Forest
NW British
Columbia,
CAN (Kitmat
and Prince
Rupert
0-80 meq
S0427m2/yr
and
0-30 meq
N/m2/yr
None
Model: Weathering rates were modeled by PROFILE and A2M
solver using empirical soil data to parameterize the models. Rates
ranged between 19 and 393 meq/m2/yr (average: 76) in the top
50 cm of soil in Prince Rupert and between 24 and 118 meq/m2/yr
(average: 57 meq/m2/yr) in Kitimat.
Williston et
al. (2016)
Be deposition
Mediterranean
holm-oak
(Quercus ilex)
forests ranging
from the typical
Mediterranean
climate to a
drier, more
seasonal
climate
Spain
Not specified None
Field Observation: In wetter forest sites, 55-65% of total base
cation deposition was wet deposition. Rainfall and net throughfall
were positively correlated for leaching for K+ and uptake for NhV
at all sites. Variable response between sites was found for Na+,
Ca2+, SO42" and CI". The authors suggest that the interplay of dry
deposition, leaching, and uptake at the canopy was different
depending on site climate and air quality characteristics.
Aquillaume
et al.
(2017)
Al = aluminum; AOSR = Athabasca oil sands region; Be = base cations; BCE = exchangeable base cations; Ca2+ = calcium ion; Fe = iron; g = gram; ha = hectare; K+ = potassium ion;
kg = kilogram; m = meter; Mg2+ = magnesium ion; N = nitrogen; Na+ = sodium ion; NH4+ = ammonium; NH4NO3 = ammonium nitrate; (NH4)2S04 = ammonium sulfate; N03" = nitrate;
S = sulfur; S04 = sulfate; yr = year.
June 2018
4-34
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
4.3.5.
Aluminum Mobilization
The 2008 ISA documented that when soil base saturation is 15 to 20% or lower,
acidifying deposition can mobilize inorganic Al, which can lead to its leaching into soil
solution and into surface waters (C'ronan and Schofield. 1990; Reuss and Johnson. 1985;
Reuss. 1983). Leaching of inorganic Al is an extremely important effect of acidifying
deposition because some forms of inorganic monomeric Al, including Al3+ and various
hydroxide species, are toxic to tree roots, fish, algae, and aquatic invertebrates
(Appendix 5 and Appendix 8). Increased concentrations of exchangeable inorganic Al in
the soil have been identified through repeated sampling in the U.S. and Europe over
periods ranging from 17 to 41 years in studies by Billett et al. (1990). Falkcngrcn-Grcrup
and Eriksson (1990). Bailey et al. (2005). and Lawrence et al. (1995).
The negative biological effects of Al mobilization are discussed in Appendix 5; in
general, Al disrupts Ca uptake by tree roots (Shortle and Smith. 1988). Substantial
evidence of this relationship has been provided through field studies (Kobe et al.. 2002;
Minocha et al.. 1997; Shortle et al.. 1997; McLaughlin and Tjoelker. 1992; Schlegel et
al.. 1992) and laboratory studies (Cronan and Grigal. 1995; Svcrdrup and Warfvinge.
1993). These studies make clear that high inorganic Al concentration in soil water can be
toxic to plant roots. The toxic response is often related to the concentration of inorganic
Al relative to the concentration of Ca, expressed as the molar ratio of Ca to inorganic Al
in soil solution. From an exhaustive literature review, Cronan and Grigal (1995)
estimated a 50% risk of adverse effects on tree growth if the molar ratio of Ca to Al in
soil solution was as low as 1.0. They estimated a 100% risk for adverse effects on growth
at a molar ratio value Ca:Al <0.2 in soil solution and minimal to no risk is thought to
occur at Ca:Al >10.
New studies on Al in soils are summarized in Table 4-6 and include an investigation of
long-term soil solution chemistry trends in Hubbard Brook Experimental Forest (Fuss et
al.. 2015). an investigation of the influence of increasing dissolved dissolved organic
matter (DOM) on toxic inorganic Al (Al,) concentration in 52 Adirondack and Hubbard
Brook watersheds (Fakhraei and Driscoll. 2015). an investigation of the effects of soil
parent material on Al (Eimil-Fraga et al.. 2016). a model comparison of three models'
predictions of soil solution Al concentrations in three monitored Swiss and German
forests (Bonten et al.. 2015). and a study showing that N addition increased Al
mobilization on grasslands in England (Horswill et al.. 2008).
Fakhraei and Driscoll (2015) evaluated the influence of increasing DOM on Al,
concentration in the Adirondacks and Hubbard Brook. The authors linked a PnET-BGC
June 2018
4-35
DRAFT: Do Not Cite or Quote

-------
1	model chemical equilibrium subroutine with an optimization algorithm. They determined
2	that accounting for the increasing concentration of DOM (as acid deposition decreases
3	and ANC increases) and its binding capacity for A1 is necessary in the model to avoid
4	substantial overestimation of available toxic A1 in waters. In an empirical study, Fuss et
5	al. (2015) found that A1 in soil water solution in the mineral horizon decreased over the
6	period of 1984-2011 (also discussed in the recovery Appendix 4.6.1).
June 2018
4-36
DRAFT: Do Not Cite or Quote

-------
Table 4-6
Aluminum mobilization.
Process/
Indicator
Type of
Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Base cation depletion Grasslands
Peak District
Not specified
Plots treated for 8
Addition: Treatments caused the Horswill et al.
Soil pH
Soil [Al], [Fe], [Mg]
National Park,

to 10 yr with 0, 35,
qrassland soils to lose 23 to 35% of their (2008)
England,

or 140 N as
total available bases and acidify by 0.2


NH4NO3
to 0.4 pH units. Al, Fe, and Mn were
mobilized and taken up by limestone
grassland forbs and were translocated
down the acid grassland soil.
Ali, DOM, DOC
Lakes
Adirondack
Long-Term
Monitoring
(ATLM) Program
lakes
Variable
None
Model: A PnET-BGC chemical
equilibrium subroutine was linked with
an optimization algorithm. Accounting for
the increasing concentration of dissolved
organic matter (DOM) as acid deposition
decreases and ANC increases, DOM's
binding capacity for Al is necessary to
avoid substantial overestimation of toxic
Ali in waters.
Fakhraei and
Driscoll (2015)
Ali
Forest and
streams
Hubbard Brook
Experimental
Forest, NH
Variable
None
Field observation/elevation gradient: Fuss et al.
Ali in soil solution in the mineral horizon (2015)
decreased over the period of
1984-2011.
Al
Forest
Three Swiss
forest monitoring
sites
Not specified None
Model: VSD modeled soil solution Al
concentrations are substantially smaller
than measured ones. However, VSD
only calculates free Al3+, whereas
measurements also include other Al
species as Al hydroxides, Al fluorides,
and Al complexed by DOM.
Bonten et al.
(2015)
Al
Spruce forest
Bechtel,
Switzerland
Not specified
None
Model: ForSAFE-modeled Al3+ was in Bonten et al.
the range of the measured soil solution (2015)
values at 20 cm depth (1990-2005)
June 2018
4-37
DRAFT: Do Not Cite or Quote

-------
Table 4-6 (Continued): Aluminum mobilization.
Process/
Indicator
Type of
Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Al, Ca and protons
Norway spruce
forest
Long-term	Not specified
monitoring site in
Germany
None
Model: SMARTml results showed that
modeled Al3+, Ca2+, and H+ agree with
measured values in soil solution
(1990-2005).
Bonten et al.
(2015)
Total Al, reactive Al
Maritime pine
forest
Northwest Spain Not specified
None
Field Observation: The soil parent
material (slate, biotitic schist, mica
schist, and granite) significantly affected
total and reactive Al in soil solution.
Total Al ranged from 17.2 to
64.2 |jmol/L. (p = 0.0006). The
concentration was significantly higher for
soil developed from mica schist
developed soil than granite and biotitic
schist.
Eimil-Fraqa et
al. (2016)
Al = aluminum; Ah = toxic inorganic Al; AOSR = Athabasca oil sands region; Be = base cations; BCE = exchangeable base cations; Ca2+ = calcium ion; Fe = iron; g = gram;
H+ = hydrogen ion; ha = hectare; K+ = potassium ion; kg = kilogram; m = meter; Mg2+ = magnesium; N = nitrogen; Na+ = sodium ion; NH4+ = ammonium; NH4N03 = ammonium
nitrate; (NH4)2S04 = ammonium sulfate; N03" = nitrate; S = sulfur; S042" = sulfate; yr = year.
June 2018
4-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
4.3.6. Nitrification and Denitrification
Nitrification is the microbial oxidation of ammonia or ammonium to form nitrite
followed by the oxidation of the nitrite to form NOs . The 2008 ISA documented that
oxidation of ammonia into nitrite is performed by two groups of organisms,
ammonia-oxidizing bacteria and ammonia-oxidizing archaea. The rate of nitrification is
controlled by numerous factors, including substrate availability (presence of NH44"),
aeration (availability of O2, often as well-drained soils with <60% soil moisture), and
acidity (pH). N addition may increase nitrification, which is often stimulated in soils with
a C:N ratio below about 20 to 25 (Ross et al.. 2004; Aber et al.. 2003). The microbial
process of autotrophic nitrification is an acidifying process, releasing 2 moles of
hydrogen ion (H+) per mole of NH4 converted to NO3 (Rcuss and Johnson. 1986). As
the N cycle becomes enriched through cumulative N addition, N becomes more abundant,
competition among organisms for N decreases, net nitrification rates often increase, and
N03 can leach from the soil (Aber et al.. 2003; Aber et al.. 1989).
Soils with a C:N ratio below about 20 to 25 are associated with stimulated mineralization,
nitrification, and cation leaching (Ross et al.. 2004; Aber et al.. 2003). This observation
makes the C:N ratio an especially useful field measurement that provides a relative index
rather than a quantitative rate of N leaching (Ross et al.. 2004). C:N ratios in the forest
floor are generally inversely related to acidifying deposition levels, although the
relationship is stronger for hardwood stands than conifer stands (Aber et al.. 2003). The
C:N ratio is a reliable and relatively straightforward measure for identifying forest
ecosystems that may be experiencing soil acidification and base leaching as a result of N
input and increased nitrification.
Denitrification is the microbial process that transforms NO3 by anaerobically reducing it
to nitrite (NO2 ). nitric oxide (NO), the greenhouse gas nitrous oxide (N2O) and N2. The
2008 ISA documented that in terrestrial ecosystems, denitrification mainly occurs in
oxygen-depleted soils (e.g., during periods of water saturation), groundwater, and
riparian zones. Soil pH has a marked effect on denitrification, with lower rates in acidic
as compared with alkaline conditions (Yamulki et al.. 1997). Soil denitrification and N2O
production and consumption are extremely variable in time and space (McClain et al..
2003).
New information published since 2008 is summarized in Table 4-7. New empirical work
along an N deposition gradient in Oregon found that nitrification increased with
increasing N deposition (Perakis and Sinkhorn. 2011); likewise, denitrification increased
with N deposition in northeastern forests (Morse et al.. 2015a). In a deposition exclusion
June 2018
4-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
study in Europe, after two decades of deposition exclusion, net nitrification and NOs
concentration in soils were not detectible, and in fact, the soil switched from a net source
of NOy to a next sink (Eickenscheidt and Brumme. 2012).
A growing body of information indicates that increased N deposition also alters the soil
microbial community (Frccdman et al.. 2013). Discussion of this is provided in
Appendix 6. Marusenko et al. (2013) explored the role of fungi in NO3 and N2O
production in soils from regions across the southwestern U.S. and found that fungi are
significant sources of N2O production in soils in semiarid grasslands and deserts. High
soil organic matter is associated with increased rates of nitrification in forest watersheds
in North America, Europe, and Japan (Mitchell. 2011). However, Russow et al. (2008)
found that soils with high soil organic matter adsorbed added NH4+, making it difficult to
determine microbial activity.
Several new syntheses evaluated N addition effects on denitrification and nitrification in
terrestrial ecosystems (Yang et al.. 2017; Bouwman et al.. 2013; Lu etak. 2011a; Liu and
Greaver. 2009). Globally, the amount of N removed from ecosystems by dentification
may be higher in terrestrial ecosystems than from groundwater or riparian zones
(Bouwman et ak. 2013); however, other older estimates have indicated more
denitrification may occur in riparian wetlands and in first-order streams than in terrestrial
ecosystems (Van Breemen et ak. 2002). although the authors acknowledged this estimate
likely underestimated terrestrial N. Liu and Greaver (2009) showed that N addition
significantly increased denitrification from all ecosystems tested (coniferous forest,
deciduous forest, tropical forest, wetland, grassland) except heathland. Among the five
chemical forms of N fertilizer added to ecosystems in studies, NO3 showed the strongest
stimulation of N2O emission (Figure 4-5). Lu et ak (2011a) further confirmed that N
addition stimulates nitrification and denitrification (Figure 4-6). Using data extracted
from 206 peer-reviewed papers, the meta-analysis showed that the largest changes caused
by N addition in the ecosystem N cycle were increased nitrification (154%), nitrous oxide
emission (134%), and denitrification (84%). In addition, Yang et ak (2017) evaluated N
cycling in five biomes in California and found a strong (r2 = 0.34) significant linear
correlation between NO3 and nitrification rates. They also found a strong negative
relationship between gross nitrification and soil C:N in forests that had soil C:N ratios
greater than 20. However, the authors noted that not all forests have such high soil C:N
ratios as those in their study that were dominated by coniferous trees. Coniferous forest
generally have higher litter C:N ratios than deciduous forests, and deciduous forests with
lower C:N ratio may also have negative relationship with C:N and nitrification.
4-40
DRAFT: Do Not Cite or Quote

-------
Table 4-7
Nitrification and denitrification.







Ambient N/S






Deposition
N/S Addition

Reference
Process Endpoint Type of Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
Nitrification and
Agricultural black
Central
Not specified
81.3 (KNOs,
Isotopic Tracer: Addition of 15N revealed
Russow et al.
denitrification
earth soils (haplic
Germany

Ca[N03]2);
denitrification of NO3" represents the main
(2008)

chernozem); two


80 kg (NH4)2S04
pathway of soil N2O release. On average,


sites: high and


76 and 54% of N2O was emitted during


normal SOM



denitrification from soils with high and






normal SOM content, respectively.






Denitrification contributed, on average, only






17 and 12% of released NO from soil with






high and normal SOM content,






respectively.

Nitrification
Spruce plantation
Hoglwald,
30 (two
None
Time series: Dynamic internal N cycle
Kreutzer et al.


Bavaria,
decades)

within the soil, driven by growth and death
(2009)


Germany


of the microbial biomass, which turns over






approximately seven-fold each year.

Denitrification
Agricultural crop,
Global
Not specified
10 to 562
Meta-analysis: Analysis of
Liu and

forest, grassland,



313 observations across all ecosystems
G re aver

wetland, tundra,



show N addition increased N2O emission
(2009)

heathland, and



by 216%.


desert





Nitrification
Agriculture and
Not specified
Mean = 105 Tg
Oto >100
Meta-analysis of 206 papers on responses
Lu et al.
Denitrification
nonagriculture

N/yr

of ecosystem N cycle in response to N
(2011a)




addition. Increases in nitrification (154%),






N2O emission (134%), and denitrification






(84%) were found. Increased N2O






emissions and N leaching under N addition






tended to export the N out of the systems






rather than benefit plant uptake over the






long term, suggesting a leaky terrestrial N






system.

June 2018
4-41
DRAFT: Do Not Cite or Quote

-------
Table 4-7 (Continued): Nitrification and denitrification.
Process Endpoint Type of Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Nitrification and
denitrification
Microbial N demand
Temperate hardwood
and conifer forests
(unfertilized)
Nine sites in the
north-central
Oregon coast
range, U.S.
Not specified None
Soil and Foliar N Gradient: As future
reductions in N deposition to polluted sites
occurs, symptoms of N saturation are most
likely to persist where soil N content
remains elevated. Temperate and
hardwood forests of the north-central
Oregon coast range showed the ratio of
netgross N mineralization and nitrification
increased along the gradient, indicating
progressive saturation of microbial N
demands at high soil N.
Perakis and
Sinkhorn
(2011)
Nitrification
Common garden
Denmark
13 to 19 for
Isotopic Tracer: Litter 815N was positively
(Callesen et

experiment, five

broadleaf
correlated with N status based on
al.. 2013)

broadleaved tree

forest; 18 to
nitrification, as well as other factors. A


species

26 for Norway
spruce
linear relationship was found between
fungal mycelia production and net
nitrification rate in lab incubations of soils
collected in the field.

Denitrification
Forested watershed
Pond Branch in
Maryland, U.S.
10 ±4 None
Isotopic Tracer: Spatial and temporal
extrapolations of measured rates suggest
that a minimum of 16-27% of atmospheric
N deposition is lost to denitrification.
Duncan et al.
(2013)
Fungal nitrification
and denitrification
Semiarid grasslands Arizona and Not specified
New Mexico
10 g soil to a
100 mL solution of
50 mmol/L
(NH4)2S04,
0.2 mol/L K2HPO4,
and 0.2 mol/L
KH2PO4, with a
pH of 7.2.
Moisture X Biocide: Fungi are significant
sources of N2O production in soils from
semiarid grasslands and deserts,
expanding evidence that fungi play a vital
role in the N cycle of arid lands.
Marusenko et
al. (2013)
June 2018
4-42
DRAFT: Do Not Cite or Quote

-------
Table 4-7 (Continued): Nitrification and denitrification.



Ambient N/S






Deposition
N/S Addition

Reference
Process Endpoint
Type of Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
Denitrification
Agricultural land and
Global
24 to 46 Tg
Not applicable
Model: N2 production from denitrification
Bouwman et

natural ecosystems

N/yr from
1900-2050

increased from 52 to 96 Tg/yr between
1900 and 2000, and N2O emissions
increased from 10 to 12 Tg N/yr. The
scenarios suggest a further increase to
142 Tg N2 and 16 Tg N20-N/yr by 2050.
Riparian buffer zones are an important
source of N2O, contributing an estimated
0.9 Tg N20-N/yr in 2000.
al. (2013)
Nitrification and
Sugar maple
Upper Michigan
15 to 20
30 kg N in the
Addition: NO3" addition to forest stands
Freedman et
denitrification
dominated northern


form of NaNC>3
across a 500-km climatic gradient
al. (2013)
Microbial community
hardwood forest


pellets delivered
decreased the abundance and richness of




to the forest floor
over the growing
season
key protein-coding genes in archaea and
bacteria responsible for N fixation,
ammonification, denitrification, and
assimilatory NO3" reduction; the same was
true for bacterial genes mediating
nitrification and dissimilatory NO3"
reduction.

Denitrification	Northern hardwood HBEF, White 6 to 8	None	Method Comparison: Both the isotopic Kulkarni et al.
forest	Mountain	tracer and gas-flow soil core method	(2014)
National Forest,	indicate that denitrification is higher and
NH	N20:N2 ratios are lower (<0.02) than
previously thought in the northern
hardwood forest and that short-term abiotic
and biotic transformations of atmospheric
N deposition to gas are significant in this
ecosystem.
June 2018
4-43
DRAFT: Do Not Cite or Quote

-------
Table 4-7 (Continued): Nitrification and denitrification.
Process Endpoint Type of Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Nitrification
Mixed hardwood
forested headwater
catchments
South-central
Ontario, Canada
Average NO3
N deposition
2.78 ± 1.22;
average NhV
N deposition
3.90 ± 1.39
Not applicable Time Series: Seasonal differences in
nitrification were largely driven by
temperature, soil moisture, and inorganic N
concentration in soil.
Annual nitrification fluxes were almost two
orders of magnitude greater than N
deposition or NO3" leaching. Nitrification
rates scaled up to annual catchment-scale
production of NO3"; the resulting fluxes are
64.9 ± 8.7 and 59.7 ±3.1 kg N/ha/yr, which
greatly exceed seasonal inputs in
deposition.
Rates of nitrification and mineralization
were similar, indicating that almost all
mineralized N is converted to NO3"
(ranging from 71 to 99%).
Casson et al.
(2014a)
Nitrification
denitrification
Deciduous and
coniferous forests
Ontario,
Canada; New
Hampshire; and
Maine
Three sites
along a
gradient: 4.5,
7, and 11
Not applicable Gradient: N gas flux increased
systematically with natural N enrichment
from soils with high nitrification rates.
N gas fluxes were linked to patterns of N
availability in forests; results do not
suggest that these fluxes respond to
increases in atmospheric N deposition at
the study sites.
Morse et al.
(2015a)
June 2018
4-44
DRAFT: Do Not Cite or Quote

-------
Table 4-7 (Continued): Nitrification and denitrification.
Process Endpoint Type of Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Nitrification
C:N
Five biome classes in California	0.6-18.4 kg
California: desert,	N/ha/yr
grassland, shrubland,
forest, wetland
Grassland soil:
76.4 kg N/ha/yr as
NPK (29:3:4);
desert soil: 60 kg
N/ha/yr as
NH4NO3;
shrubland soil:
60 kg N/ha/yr as
NH4NO3
N Addition: Across biomes, a positive
correlation of gross nitrification to soil NO3"
(r2= 0.34) and a negative correlation to soil
C:N (r2 = 0.31) was observed.
No correlation was found between gross N
mineralization and nitrification. Deserts had
the lowest gross N mineralization rates and
exhibited similar nitrification rates to the
shrublands and grasslands. Only 15% of
NH4+ produced was nitrified in the forests
compared to 47 to 86% in the other
biomes. This suggests that NH4+
production rates did not limit nitrifiers in the
forests, but rather, that another factor
limited nitrifier activity.
Yang et al.
(2017)
Ca(N03)2 = calcium nitrate; g = gram; HBEF = Hubbard Brook Experimental Forest; K2HP04 = dipotassium phosphate; KH2P04 = monopotassium phosphate; KN03 = potassium
nitrate; L = liter; mL = milliliter; mmol = millimole; mol = mole; N = nitrogen; N2 = molecular nitrogen; N20 = nitrous oxide; N20-N = nitrogen from nitrous oxide; NaN03 = sodium
nitrate; NH4+ = ammonium; (NH4)2S04 = ammonium sulfate; NO = nitric oxide; N03" = nitrate; SOM = soil organic matter; Tg = teragram yr = year.
June 2018
4-45
DRAFT: Do Not Cite or Quote

-------
(a) aii
Mean (156)
i—*-
(b)
Agriculture (aerobic) (3S)
Agriculture (artaercttic:) (1B) I—•—I
Ooniterous (33)
Deciduous (15)	f-
Tropicai forest (.11)
Wellarvd (19)	j I-
Grassland (16)	| t-
Heathland (3)	(-
(c) N form
NH,NO= (44)
NH/ t2&)
NO," (24}
Urea (32)
UAN (23)
(d) Experimental length
Short term (61)
Long term (90)
(e) N addition level
<65 (36)
55-150(29>
>150 (51)
I—•-

3 4 5 6 7 6 9 1S 16 20
Response ratio
N = nitrogen; NH4N03 = ammonium nitrate; NH4+ = ammonium; N03 = nitrate; UAN = urea and ammonium nitrate fertilizer.
The data are expressed as the mean response ratio with 95% confident intervals. The numbers of studies included are indicated in
parentheses.
Source: Liu and Greaver (2009).
Figure 4-5 Effects of nitrogen addition on biogenic nitrous oxide emission.
June 2018
4-46
DRAFT: Do Not Cite or Quote

-------
3
Leaf N
Aboveg round plant N
Belowground plant N
Litter N
Microbial N
Organic horizon N
Soil N
DON
Soil inorganic N
N mineralization
Immobilization
Nitrification
Denitrification
Leaching ¦
0.0 0.4 0.8 1.2 1.6 2.0
Weighted response ratio (RRJ
N = nitrogen; N20 = nitrous oxide; RR = response ratio.
Bars represent RR++ ± SE. The vertical line is drawn at logeRR = 0. The sample size for each variable is shown next to the bar.
Source: Lu et al. (2011a).
Figure 4-6 The weighted response ratio for the responses to nitrogen
addition for fluxes and pools related to the ecosystem nitrogen
cycle in agricultural (open bars) and nonagricultural (closed bars)
ecosystems.
4.3.7. Decomposition
Decomposition is a general term that refers to the breakdown of organic matter.
(Schlesinger. 1997). Decomposition is an important part of N and C cycling that can be
altered by N deposition. Decomposition rates correlate with ratios of C:N, lignin:N, and
June 2018
4-47
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
lignin:cellulose in litter, all which may be altered by N deposition. The addition of N can
stimulate the decomposition of labile compounds that degrade during the initial stages of
decomposition, but added N can suppress the decomposition of more recalcitrant
material. Evidence for this is widespread in forests, but has not yet been well documented
in grasslands and other ecosystems. Since 2008, there are new addition studies and
meta-analyses to better understand the mechanisms and response trends.
The 2008 ISA documented that the soil microbial community (bacteria and fungi) are the
main decomposers of organic matter. Both the microbial community composition and
microbial enzyme activity can dynamically respond to shifts in inorganic nutrient and
substrate availability (Compton et al.. 2004; Carreiro et al.. 2000); the shift reflects the
nutrient and energy limitation of the microbial community. Litter decay rates are also
well established to correlate with ratios of C:N, lignin:N, or lignin:cellulose in litter
(Hobbic. 2008; Aerts. 1997; Melillo et al.. 1982). These chemical traits are strong
predictors of litter decay, accounting for over 73% of the variation in litter decomposition
rates worldwide (Zhang et al.. 2008).
Traditionally, carbon dioxide (CO2) is measured as an indication of soil respiration and a
proxy for decomposition. Since the 2008 ISA, analysis of microbial enzymes and genes
in the soil have been used to identify microbial activity. Litter decomposition is an
enzymatically complex process mediated by a diverse assemblage of microorganisms that
includes decomposition from both autotrophic (root) biomass and heterotrophic
(microbial) activity. More information is necessary to attribute the autotrophic or
heterotrophic source apportionment. Numerous studies have been published since 2008
describing how N addition affects the decomposition of organic C and N (Table 4-8).
4-48
DRAFT: Do Not Cite or Quote

-------
Table 4-8 Decomposition.
Process/
Type of

Deposition
Addition

Reference
Indicator
Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
Decomposition
Sugar maple
Bear Brook

1,800 eq/ha/yr
N + S Addition: Caused increased N concentration in
Hunt et al.

hardwood forest
Watershed in

of (NH4)2S04
leaves and faster short-term decomposition.
(2008)


eastern Maine




Enzyme activity
Eight annual
Controlled
2.0 and 44.0

Growth Chamber: N deposition increased soil enzyme
Mannina et al.
Decomposition
herb species
ecosystem (litter


activity known to breakdown cellulose (cellobiosidase,
(2008)

bags placed in


p-glucosidase and (3-xylosidase).



annual






herb-based






microcosm






ecosystems)




Enzyme activity
Northern
Michigan
6.8 to 11.8
30 NOs"
N Addition: N addition altered rates of organic matter
Preqitzer et
Decomposition
hardwood forest



decomposition by suppressing the soil enzymes
al. (2008)




responsible for litter degradation when litter has a higher N






concentration. This causes an increase in surface soil C






storage.

N Addition: Generally stimulated activities of cellulose Keeler et al.
degrading enzymes in litter and soil, but had no effect on (2009)
lignin degrading enzyme activity. N addition had a negative
or neutral effect on litter and SOM decomposition in the
same sites, with no correspondence between effects of N
on enzyme activity and decomposition across sites.
Enzyme activity
Northern
hardwood forest
Catskill
Mountains of
New York, U.S.
50 (NH4NO3)
N Addition: Identified that patterns in microbial community
structure and function were more strongly influenced by
tree species than by fertilization.
Weand et al.
(2010)
Soil respiration
Decomposition
Forest
Global
Varied
Meta-Analysis: N additions decreased root respiration,
heterotrophic respiration, and soil CO2 but had no effect on
litter decomposition.
Janssens et
al. (2010)
Enzyme activity Forests	Central	Not	100 (NH4NO3)
Decomposition grasslands	Minnesota	specified
June 2018
4-49
DRAFT: Do Not Cite or Quote

-------
Table 4-8 (Continued): Decomposition.
Process/
Indicator
Type of
Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Soil respiration
Temperate
broadleaf and
mixed forest,
temperate
conifer forest,
boreal forest,
tropical forest,
grassland,
wetland, tundra,
desert, and arctic
Global
Varied
Meta-Analysis: N additions decreased heterotrophic
respiration but had no effect on total soil respiration.
Liu and
G re aver
(2010)
Enzyme activity
Decomposition
Forest and
grasslands
U.S.
None
None
Synthesis: Data from 28 ecosystems show resources to
produce the enzymes phenol oxidase and
(3-1,4-glucosidase are uncoupled. This indicates that the
increasing recalcitrance of organic matter decreases C and
nutrient availability and slows microbial growth.
Sinsabauqh
and Shah
(2011)
Enzyme activity
Mature black
Central Alaska
Not
Years
N Addition: N fertilization may alter decomposer
Talbot and
Decomposition
spruce forest in

specified
2009 = 200
community structure by favoring a shift toward cellulose-
Treseder
upland boreal


(NH4NO3)
and mineral-N users. Cellulose degrading microbes
(2012)

ecosystem


Years
(decomposers) were competitively dominant under N





2010 = 100
fertilization.





(NH4NO3)


Enzyme activity
Decomposition
Sugar maple
forests
Michigan
Not
specified
30
Meta-Analysis: N addition increased cellulose
decomposition by 9% and decreases lignin decomposition
rates by 30%. Overall, N increases the amount of litter
mass entering the humus pool and leads to increases in
soil C storage under experimental N deposition.
Whittinqhill et
al. (2012)
Enzyme activity Northern
Decomposition hardwood forest
Minnesota
N Addition: Accelerated the initial decomposition rate.
Faster initial decomposition rates corresponded to higher
activity of polysaccharide-degrading enzymes and greater
relative abundances of Gram-negative and Gram-positive
bacteria. Later in decomposition, externally supplied N
slowed decomposition, increasing the fraction of slowly
decomposing litter, reducing lignin-degrading enzyme
activity, and relative abundances of Gram-negative and
Gram-positive bacteria.
Hobbie et al.
(2012)
June 2018
4-50
DRAFT: Do Not Cite or Quote

-------
Table 4-8 (Continued): Decomposition.
Process/
Type of

Deposition
Addition

Reference
Indicator
Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
Decomposition
Deciduous
Catskill
9.0
50 mg N/ha/yr
N Addition: N addition caused decrease in mineralization
Lovett et al.

forests
Mountains of

N (NH4NO3)
and nitrification and an increase in forest floor C pools and
(2013)


southeastern


C:N, indicating that N addition increased C sequestration in



New York, U.S.


the organic horizons of the soil, most significantly in






hemlock plots.

Enzymes	Northern	Upper Michigan 15 to 20 30	N Addition: Caused slower organic matter decay and Freedman et
Decomposition hardwood forest	altered microbial community composition and function. al. (2013)
Observed a decrease in the abundance and richness of
key protein-coding genes in archaea and bacteria
responsible for N fixation, ammonification, denitrification,
and assimilatory NO3" reduction; the same was true for
bacterial genes mediating nitrification and dissimilatory
NO3" reduction.
N Addition: (1) significantly altered the composition of Eisenlord et
actinobacterial and fungal genes mediating plant and	al. (2013)
fungal cell wall depolymerization; (2) significantly
decreased the richness and diversity of genes involved in
the depolymerization of starch (-12%), hemicellulose
(-16%), cellulose (-16%), chitin (-15%), and lignin (-16%);
and (3) resulted in small changes in community
composition (25% difference in fungi; 18% in
actinobacteria).
Enzymes
Northern
Upper Michigan
15 to 20
30 as six equal
N Addition: Observed that atmospheric N deposition
Freedman
Decomposition
hardwood forest


applications of
increases saprotrophic bacterial laccase-like multicopper
and Zak



NaN03 pellets
delivered to the
forest floor over
the growing
season
oxidases (LMCOs). These results suggest a plausible
mechanism by which anthropogenic N deposition has
reduced decomposition, increased soil C storage, and
accelerated phenolic DOC production.
(2014)
Decomposition
Forest
Switzerland
Not
7 (NH4NO3)
N Addition: Promoted the production of new fungal
Grieoentroa
fungal residue


specified
70 (NH4NO3)
residues but slowed the decomposition of old residues in
et al. (2014)




forest soil fractions. Preservation of old microbial residues
could be due to decreased N limitation of microorganisms
and therefore a reduced dependence on organic N
sources.

Enzymes	Sugar maple Michigan	5.8 to 7.3 30
Decomposition forests
June 2018
4-51
DRAFT: Do Not Cite or Quote

-------
Table 4-8 (Continued): Decomposition.
Process/
Indicator
Type of
Ecosystem
Region
Deposition
kg/ha/yr
Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Enzyme activity Northern	Bear Brook
hardwood and Watershed
softwood forest (BBWM), ME,
U.S.
25.5
N Addition: After 22 yr of N addition, N enrichment had Mineau et al.
little effect on microbial enzyme activity in terrestrial	(2014)
compartments, even across varying degrees of organic
matter recalcitrance.
Soil respiration Multiple biomes Global
Varied
Meta-Analysis: N addition significantly increased soil
respiration by 2.0% across all biomes but decreased
respiration by 1.44% in forests and increased it by 7.84 and
12.4% in grasslands and croplands, respectively (p < 0.05).
The response ratios of soil respiration to N addition were
positively correlated with mean annual temperature (MAT),
most significant when MAT was less than 15°C. N addition
largely altered root and microbial biomass and soil C
content, which are likely the mechanisms behind the
altered soil respiration.
Zhou et al.
(2014a)
Decomposition Grassland
Minnesota,
Nebraska, Iowa,
Kansas;
Colorado
3.1 to M
100
N Addition: Decreased microbial respiration of OM by as Riggs et al.
much as 29% relative to control plots, and consequently, (2015)
decreased C loss from this pool.
Decomposition Forest
Great Lakes
region
6.8 to 11.8 30 as NaN03
N Addition Fine root biochemistry was less responsive
than leaf litter to long-term simulated N deposition. Fine
roots were the dominant source of difficult-to-decompose
plant C fractions entering the soil.
When combined with litter production, simulated N
deposition increased N flux through leaf litter by an average
of 29%, but did not affect fine root N flux.
Xia et al.
(2015)
Enzyme activity Subalpine forest
Loch Vale
watershed
(LVWS) is
located in Rocky
Mountain
National
Park (RMNP) on
the eastern edge
of the Colorado
Front Range.
3-4 kg 25 kg	N Addition: There were no changes in the C degrading
N/ha/yr NhUNOs/ha/yr enzyme activity in response to fertilization, while the N
(wet)	(1996-present) degrading enzyme activity was enhanced with elevated
activity of leucine aminopeptidase (LAP) and marginally
significant increase in N-acetyl-p-glucosaminidase (NAG)
with N fertilization.
Boot et al.
(2016)
June 2018
4-52
DRAFT: Do Not Cite or Quote

-------
Table 4-8 (Continued): Decomposition.
Process/
Type of

Deposition
Addition

Reference
Indicator
Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
Enzyme activity
Northern
Michigan
5.9-7.4 kg
30 kg
N Addition: This long-term experiment found N deposition
Freedman et

hardwood forest

N/ha/yr (not
NaN03/ha/yr
decreased the activity reduced the activity of extracellular
al. (2016)



specified if
(beginning in
enzymes mediating plant cell wall decay.




wet, dry, or
1994)





total)



Soil respiration
Northern
White Mountain
8 kg N/ha/yr
30 kg N/
N Addition: The greatest reduction in soil respiration on N
Kana et al.

hardwood forest
National Forest,
(wet + dry)
ha/yr
and N + P fertilized plots occurred on the sites with lowest
(2016)


NH; Bartlett,
(20th
(NH4NO3),
pretreatment soil N mineralization and litterfall N flux.



Hubbard Brook,
century
10 kg P/ha/yr
Nutrient additions did not significantly affect either fine root



and Jeffers Brook
average at
(NaH2P04), or
turnover (minirhizotrons) or microbial respiration (laboratory



Forests
HBEF)
N + P
incubations).





(treatments in






Year 2011)


Decomposition
Restored
Prairie Invasion
-1/3 ofN
5 g N/m2/yr
N Addition: Increasing rainfall variability and N addition
Schuster

tallgrass prairie
and Climate
added
urea. "Seasonal
can stimulate litter decomposition in tall grass prairie.
(2016)


Experiment

maximum




(PRICLE)

5-day




Loveland, CO

cumulative






rainfall"






increased by 33






(2012) and 9%






(2013)


Decomposition
All ecosystems
Global
Not
Not specified
Meta-Analysis: 198 peer-reviewed journal articles found N
Yue et al.



specified

addition did not significantly alter litter decomposition, soil
(2016)





respiration (except for wetlands, +28.26%), and microbial






respiration (except for forests, +9.08).

June 2018
4-53
DRAFT: Do Not Cite or Quote

-------
Table 4-8 (Continued): Decomposition.
Process/
Type of

Deposition
Addition

Reference
Indicator
Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
Decomposition
Mixed hardwood
Chronic Nitrogen

50-150
Changes in litter decay by was generally lower in the N
van Dieoen et

forest dominated
Amendment


treatment microbes compared to control microbes for the
al. (2017)

by black and red
Study (CNAS)


same species, a response not readily reversed when N


oak (Quercus
located at the


microbial isolates were grown in low N environments.


velutina and Q.
Harvard Forest


Changes in fungal behaviors accompany and perhaps drive


rubra)
Long-Term 57


previously observed N induced shifts in fungal diversity,



Ecological


community composition, and litter decay dynamics.



Research (LTER)






site in






Petersham, MA,






USA




Decomposition
Forest
Scotland
14-16 kg
1.18 g 15N in
Litter decomposition is a larger source of N for trees than
Nair et al.



N/ha/yr
4m plot
simulated N deposition.
(2017)
BBWM = Bear Brook Watershed; C = carbon; DOC = dissolved organic carbon; ha = hectare; kg = kilograms; LMCO = Laccase-like multicopper oxidase; mg = milligrams; N = nitrogen;
15N = nitrogen-15; NaN03 = sodium nitrate; NH4N03 = ammonium nitrate; (NH4)2S04 = ammonium sulfate; N03" = nitrate; 15N03" = nitrogen-15-labeled nitrate; OM = organic matter;
S = sulfur; SOM = soil organic matter; yr = year.
June 2018
4-54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
The body of knowledge published since the 2008 ISA indicates the effects of N on
decomposition rates are inconsistent among studies. New field work supported the
concept that N deposition suppresses decomposition (Riggs et al.. 2015; Zak et al.. 2008).
Likewise, Kang et al. (2016). in a large-scale N addition study of 13 northern hardwood
forests in the U.S., supported the findings that N addition decreases forest respiration and
further identified the greatest reduction in soil respiration on N fertilized plots occurred
on the sites with lowest pretreatment soil N mineralization and litterfall N flux (Kang et
al.. 2016). In contrast, meta-analysis that evaluated the central tendencies of N addition
on total soil respiration in forest biomes reported inconsistent results, including an
increase (Yuc et al.. 2016). decrease (Zhou et al.. 2014a). and no effect (Liu and Grcavcr.
2010). Meta-analyses that looked for the central tendencies of terrestrial soil respiration
to N addition also have inconsistent results; two meta-analyses found that N did not alter
total soil respiration in terrestrial soils (Yue et al.. 2016; Liu and Greaver. 2010). another
identified an increase (Zhou et al.. 2014a). and only one identified a decrease (Janssens et
al.. 2010). The different results in the meta-analyses reflect slightly different selection
criteria the authors used to determine which studies to include in the analyses. Among
biomes, differences in soil respiration may result largely from stimulation of autotrophic
respiration by N addition to croplands and grasslands compared with no significant
change for forests, and a simultaneous decline in heterotrophic/microbial respiration in
most biomes I (Zhou et al.. 2014a; Liu and Greaver. 2010). with the exception of
croplands, tropical forests, and boreal forests].
One proposed mechanism for reduced microbial decay under increased N deposition is a
shift in the species composition of the microbial community with the consequence of
decreased lignin decomposition. Lignin is an organic polymer, particularly important in
cell wall formation of vascular plants. Under N addition, the microbial community would
shift from basidiomycete fungal activity, some of which oxidize lignin in plant detritus
and polyphenols in SOM to CO2, to more metabolism by bacteria and ascomycete fungi
(cellulose degraders), which only partially oxidize these organic substrates (Freedman
and Zak. 2014; Zak etal.. 2011). There is new evidence to support this theory in
northeastern hardwood forest in the U.S. where N enrichment increases the species
richness and diversity of ascomycetes generally (cellulose degraders) (Morrison et al..
2016). with evidence that observed changes in decay abilities were not readily reversed
when N isolates were grown in control environments, indicating that the fungal
community may not recover quickly following the cessation of N enrichment (van
Diepen et al.. 2017) (see Appendix 6).
There is new evidence to support that lignolytic enzyme activity decreases under N
addition (Freedman et al.. 2016; Keeler et al.. 2009; Manning et al.. 2008). This trend is
also supported by a meta-analysis by Whittinghill et al. (2012). who found that N
4-55
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
addition decreased lignin decomposition rates by 30% and increased cellulose
decomposition by 9%. Fine roots are likely the dominant source of
difficult-to-decompose plant carbon fractions entering the soil, and this pattern appears to
be widespread in boreal and temperate forests (Xia et al.. 2015).
There are cases in which N addition causes no change in enzyme activity of microbes; for
example, the hardwood and softwood forests growing in the Bear Brook Watershed in
Maine (Mincau et al.. 2014) and C degrading enzymes in a subalpine forest (Boot et al..
2016). Another study in the Catskill Mountains of New York State found that patterns in
microbial community structure and function (detected by enzyme activity) were more
strongly influenced by the tree species present than by fertilization (Wcand et al.. 2010).
Microbial decomposition response to N addition may also change through time. The 2008
ISA documented an N addition study lasting longer than 2 years that indicated that a shift
can occur from stimulation to depression of decomposition overtime (Knorr et al.. 2005).
In a study looking at changes through time of the organic laver.Hobbie et al. (2012)
suggest N deposition in forest ecosystems may decrease mean residence times of active
fractions in fresh litter, while increasing those of more slowly decomposing fractions,
including more processed litter. In boreal forests, Talbot and Treseder (2012) found a
transition over time from competition among decomposers to high cellulase activity and
suppressed lignin loss under N fertilization. The trend suggests that, in N limited systems,
N fertilization may alter decomposer community structure by favoring a shift toward
cellulose- and mineral-N users.
There is also new evidence that experimental N deposition significantly decreased the
richness and diversity of microbial genes involved in the depolymerization of starch
(12%), hemicellulose (16%), cellulose (16%), chitin (15%), and lignin (16%). (Eisenlord
et al.. 2013).
4.3.8. Nitrogen Mineralization
Mineralization refers to processes that release carbon as CO2 and nutrients in inorganic
form. Nitrogen mineralization is the process by which organic N is converted to
plant-available inorganic forms. The 2008 ISA documented that the rate of mineralization
may be influenced by numerous factors including C:N of soil organic matter, soil pH, and
the microbial community. N mineralization has been shown to increase with increasing N
addition (Aberetal.. 1998). often up to 1.6 times the control (Gundersen et al.. 1998).
New publications (Table 4-9) support that soil N mineralization increases with N addition
across terrestrial ecosystems (Lu etal.. 2011a). specifically in temperate forests (Nave et
June 2018
4-56
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
al.. 2009a). likely due to increases in dissolved organic nitrogen (DON), the size of the N
pool in the soil, and decreases in C:N ratios. The forest floor responds differently than the
deeper mineral soil layer I Figure 4-7; (Lovctt et al.. 2013; Nave et al.. 2009a)l.
In two forested headwater catchments in Ontario, Casson et al. (2014a) found that N
mineralization and nitrification rates are similar, indicating that almost all mineralized N
is converted to NOs (ranging from 71 to 99%) in catchments they studied. While Bade et
al. (2015) found in old-growth spruce forest, lower N mineralization occurred in the more
open patches than the closed ones. Possible reasons were reduced litter supply and lower
canopy N interception in gaps in this forest under exposure to high N deposition. Further
studies in other old-growth forests are needed to better understand the mechanisms
causing long-term change in N cycling with forest development.
In desert shrublands, Rao et al. (2009) found that N deposition may increase production
and/or alter litter C:N ratios that increase soil C. There was an inverse relationship
between the C:N ratio and total N mineralized, yet a positive relationship between
organic C and total N with mineralization. These results indicate that microbial activity in
low productivity arid land soils is primarily limited by C and secondarily limited by N.
4-57
DRAFT: Do Not Cite or Quote

-------
Table 4-9 Nitrogen mineralization.



Ambient N/S



Process
Type of

Deposition
N/S Addition

Reference
Endpoint
Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
Mineralization
Semiarid
Southern
Up to 35-45
Two
N is mineralized at a faster rate from grass litter
Sirulnik et al.


California,
forN
applications
enriched in N as a result of N deposition
(2007a)


U.S.

per year of






30 kg N


Mineralization
Semiarid
Southern
Up to 35-45
Two
When significantly lower microbial N was reported in
Sirulnik et al.


California,
forN
applications
in fertilized plots it corresponded to faster net N
(2007b)


U.S.

per year of
mineralization. When higher microbial N was





30 kg N
observed in fertilized plots it corresponded to net






immobilization (though not significantly faster), thus






corroborating evidence that the microbial community






was taking up more N in fertilized plots. These rates






may have corresponded to C availability, which was






not measured.

Mineralization
Mixed hardwood
Adirondack
Not specified
None
Ca Gradient: The exchangeable Ca coupled with
Paae and Mitchell
soil [N]
stands
Mountains,


soil moisture, soil organic matter, and ambient
(2008)
soil [Ca]

NY, U.S.


temperature accounted for 61% of the variability in





extractable inorganic N across 11 sites. The






influence ofCa on soil inorganic N may be through






interactions between soil Ca concentrations and






species composition, which in turn affect the quality






of litter available for N mineralization.

Decomposition
Spruce forest
Germany
8.5

N Exclusion: Some N cycling enzymes increased
Enowashu et al.
Enzyme




activities, whereas others decreased under reduced
(2009)
activities




N treatment.

N mineralization
Desert
Joshua Tree
2.7 to 14.4

Calculated soil N from deposition was directly
Rao et al. (2009)


National Park


correlated with measured soil C and N and






decreasing C:N ratios.

June 2018
4-58
DRAFT: Do Not Cite or Quote

-------
Table 4-9 (Continued): Nitrogen mineralization.
Process
Endpoint
Type of
Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
N mineralization
Temperate
forests.
Northeastern
U.S.
Varies
Varies
Meta-Analysis: Overall, N inputs increased soil C
(+7.7%) and N mineralization (+62%), while
decreasing C:N (-4.9%).
Nave et al. (2009a)
N mineralization
Forest


50 (6 yr of N
addition)
N Addition: There was a significant decline in
potential N mineralization and nitrification rates in
the mineral horizon but not in the forest floor.
Lovettetal. (2013)
N mineralization
Terrestrial
ecosystems

Varies
Varies
Meta-Analysis: of 206 papers on responses of
ecosystem N cycle caused by N addition. N addition
increased mineralization rates.
Lu et al. (2011a)
N mineralization
Douglas fir forest
Oregon coast
range
2.0
None
Aboveground N uptake by plants increased with net
N mineralization, peaking at 35 kg N/ha/yr.
Perakis and
Sinkhorn (2011)
Net N
Mixed hardwood
Muskoka-
NO3-
Not applicable In all seasons, rates of nitrification were similar to
Casson et al.
mineralization
forested
Haliburton
deposition
rates of total mineralization, indicating that almost all
(2014a)
and nitrification
headwater
district of
2.78 ± 1.22;
mineralized N is converted to NO3" in both soil types

rates/NH4+ N,
catchments
south-central
NH4+
(ranging from 71 to 99%).

soil NO3",

Ontario,
deposition


stream NO3"

Canada
3.90 ± 1.39


N
Natural
Harz National
27 (open
Not applicable Net N mineralization (and ammonification) rates Bade et al. (2015)
mineralization,
(unmanaged)
Park in
areas)
were higher in the closed stands of the optimum and
ammonifi cation,
old-growth
central
47 (closed
overmature stages than in the more open decay and
nitrification/N03"
Norway spruce
forest
Germany
forest)
regeneration stages. Only a small proportion of
NH4+ was oxidized to NO3" in the acidic soils.
Net litter
mineralization
Forest
Scotland
14-16 kg
N/ha/yr
1.18 g 15N
per 4m plot
Three times as much 15N was retained in the O and
A soil layers when N was derived from litter
decomposition than from mineral N additions.
Nairetal. (2017)
C = carbon; Ca = calcium; ha = hectare; kg = kilogram; N = nitrogen; NH4+ = ammonium; N03 = nitrate; yr = year.
June 2018
4-59
DRAFT: Do Not Cite or Quote

-------
Sol! C	C/N	Nmin
Forest floor
Overall
٦ N-fix
N5
T>
^ N-fert
'o
D
J N-dep
Mineral soil
Overall
= N-fix
"5
T3
to
z N-fert
O
<0
o
^ N-dep
-40 -20 0 20 40 60 -20 -15 -10 -5 0 5	0 100 400 600
Change with N inputs (%)
C = carbon; C/N = carbon-to-nitrogen ratio: N-dep = nitrogen deposition; N-fert = nitrogen-fertilization; N-fix = nitrogen-fixing
vegetation; Nmin = nitrogen mineralization.
Points are means ± bootstrapped 95% confidence intervals, with number of studies in parentheses. Groups with confidence
intervals overlapping the dotted reference line (0% change) show no significant effects of N addition. Soil carbon responses to
nitrogen inputs are shown separately for pool sizes (filled symbols) and concentrations (open symbols). For each response
parameter, the effects of nitrogen inputs are shown overall, and for individual modes of nitrogen addition, including the
establishment of nitrogen-fixing vegetation, large-dose nitrogen-fertilization, and simulated chronic nitrogen deposition. Note that
points plotted in panels C and F have 95% confidence intervals with upper-bounds beyond the x-axis maximum.
Source: Nave et al. (2009a).
Figure 4-7 Effects of nitrogen inputs on soil carbon, carbon:nitrogen ratio,
and minimum nitrogen in forest floors (panels A-C) and mineral
soils (panels D-F).
j (87) D
j-o- (115)
!-*- (35)
—O	(36)
j	•	(22)
4t>- (16)
"j* (30)
-d-(61)
I	I	t	1	I	I
(86)
(8)
(45)
(33)
B
i	i	i	i	i
(121)
E


(50)



(25)


(46)
—i	1	r

(81)
(37)
(42)
(2)


(80)
(51)
—A	
(24)
(5)
	r-

June 2018
4-60
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
4.3.9.
Dissolved Organic Carbon Leaching
The acid-base characteristics of dissolved organic matter (DOM) are an important part of
understanding the recovery potential for soils, lakes, and streams impaired by acid
deposition. The composition of DOM includes fulvic and humic acids, carboxylic acids,
and amino acids; dissolved organic carbon (DOC) includes these organic acids. Here we
focus on the DOC component of DOM. The many carboxylic groups of DOC make it
chemically interact like a weak acid; therefore, DOC content may affect pH levels. In
addition, Fakhraei and Driscoll (2015) emphasize the importance of predicting accurately
the acid-base properties of recovering surface waters because, for example, the acidic
components of DOC act as hosts for binding trace metals like toxic Al (for additional
discussion on Al and DOM see Appendix 4.3.5).
In recent years, the DOC of many lakes and streams has risen, with the likely source
being the soils in the adjacent terrestrial watershed. However, the mechanism causing the
observed increase is unclear. The increase may be due to a combination of soil recovery
from acidification, changes in climate (e.g., temperature and precipitation), and N
deposition, among other mechanisms (for reviews see (Evans et al.. 2005; Kalbitz et al..
2000). New studies in the literature have investigated soil DOC response to acidification
(N + S deposition) and/or looked at the effects of N addition; these studies will be the
focus of the following discussion (Table 4-10). For a discussion of DOC in surface water
see Appendix 7.
June 2018
4-61
DRAFT: Do Not Cite or Quote

-------
Table 4-10 Terrestrial dissolved organic carbon (DOC) leaching.


Ambient N/S



Type of

Deposition
N/S Addition

Reference
Process Ecosystem
Region
kg/ha/yr
kg/ha/yr
Effect of Deposition
HERO ID
DOC Bog, heathland,
Synthesis of
Ranged from
Ranged from 10
Synthesis: N had inconsistent effect on
Evans et al. (2008)
grassland,
17 addition
5-16 N dep
to 150 N
DOC; however, the form of N applied

broadleaf and
experiments in

(chemical forms
indicates nonacidifying forms of N tend to

coniferous forest
the northeastern

for N varied)
increase DOC concentrations.


U.S. and





northern Europe




DOC Temperate	410 observations Not specified
mixed and	globally
conifer forests,
boreal forests,
grasslands,
tropical forests,
arctic, wetlands,
desert, and
tundra
DOC Norway spruce Southern	Not specified Not specified Model: The Stockholm Humic Model was Lofqren et al. (2010)
(Picea abies) Sweden	used to model DOC solubility in soil water
forests	and predicted that DOC trends could vary
between positive and negative depending
on changes in pH, ionic strength, and soil Al
pools.
10 to 650	Meta-Analysis: N addition increases short- Liu and Greaver (2010)
term belowground C storage by increasing
C content of the organic layer. N addition;
increased DOC concentration (+18%); and
increased C content of the organic soil layer
(+17%) but not the mineral soil layer.
June 2018
4-62
DRAFT: Do Not Cite or Quote

-------
Table 4-10 (Continued): Terrestrial dissolved organic carbon (DOC) leaching.
Process
Type of
Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
DOC
Beech forest
stands on
calcareous soils
Swiss Jura
mountain range
Not specified 5.5 (NH4NO3)
N Addition: Leaching of DOC from the litter
layer was not affected by N additions, but
DOC fluxes from the mineral soils at 5- and
10-cm depth were significantly reduced by
17%. 13C indicated that litter-derived C
contributed less than 15% of the DOC flux
from the mineral soil, with N additions not
affecting this fraction. Hence, the
suppressed DOC fluxes from the mineral
soil at higher N inputs can be attributed to
reduced mobilization of nonlitter derived
"older" DOC.
Haaedorn et al. (2012)
DOC
None—
theoretical
None—
theoretical
Not specified Not specified
N Addition: Increases mineralization
causing the pool of bioavailable DOC to
decrease. Consequently, relatively less
bioavailable DOC remains for NhU"1"
assimilation and immobilization in microbial
biomass, leaving more NhV for nitrifiers. As
a result, internal NO3" production increased.
The higher bacterial demand for DOC under
elevated availability of N and electron
acceptors comes into conflict with
increasing chemical suppression of DOC
solubility and bioavailability in the
progressively acidifying soils and finally
results in the C limitation of microbial
metabolism.
Kopacek et al. (2013)
DOC Northern	Upper Michigan 15 to 20 kg
hardwood forest	N/ha/yr
30 (as six equal
applications of
NaN03 pellets
delivered to the
forest floor over
the growing
season)
N Addition: N addition accelerated phenolic
DOC production.
Freedman and Zak (2014)
June 2018
4-63
DRAFT: Do Not Cite or Quote

-------
Table 4-10 (Continued): Terrestrial dissolved organic carbon (DOC) leaching.
Process
Type of
Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
DOC Forest and	Hubbard Brook Variable
streams	Experimental
Forest, NH
None
Field Observation: Concentrations of DOC
showed statistically significant declines in
the Oa soil solutions of all three elevation
zones, and a subset of the Bs soil solutions
over the period of 1984-2011.
Fuss etal. (2015)
DCO Lakes	Adirondack Variable	None	Field observation: increasing DOC in 29 of Driscoll et al. (2016)
Long-Term	48 lakes as lake pH increased.
Monitoring
(ATLM) Program
lakes at
Huntington
Forest and
Whiteface
Mountain,
Adirondack
Mountains, NY
DOC Forest and	U.K.	Mean 1993 to None
grassland	2010
gradients of
deposition:
S: 44 to
86 meq/m2/yr;
N: 40 to
90 meq/m2/yr;
CI: 94 to
306 meq/m2/yr
Modeling: Using the MADOC model, the Sawicka et al. (2017)
acidifying effect of S deposition was the
predominant control on the observed soil
water DOC trends. The relative importance
of S and N loading depended on soil
sensitivity to acidification, and on N
limitation.
13C = carbon-13; C = carbon; DOC = dissolved organic carbon; ha = hectare; kg = kilogram; N = nitrogen; NaN03 = sodium nitrate; NH4+ = ammonium; NH4N03 = ammonium nitrate;
N03" = nitrate; S = sulfur; yr = year.
June 2018
4-64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
A number of studies suggest that recovery from acidification is driving the elevated
levels of DOC. An inverse relationship between mineral and organic acid export to
surface waters from soils was first proposed in the 1980s (Krug and Frink. 1983). Support
for this theory is found in a range of laboratory experiments in which increases in both
acidity and ionic strength (associated with a high SOr loading) have been shown to
reduce soil solution DOC (Kalbitz et al.. 2000). Evans et al. (2008) addressed whether
elevated N deposition or recovery from acidficiation associated with decreasing S cause
increased DOC loss from upland soil. Their analysis of a large experimental data set
including 12 sites in the U.S. and northern Europe showed that although the response to
N addition is inconsistent, DOC concentrations responded predictably to the chemical
form of N added. DOC concentrations increased with NaNOs additions or gaseous NH3
exposure, and decreased with most NH4+ salt additions. The authors cite the effect of the
chemical form of N on acidity as a plausible mechanism and further conclude that their
evaluation does not provide clear support for the role of N deposition as the sole driver of
rising DOC, but is consistent with an acidity-change mechanism. Evans et al. (2008) also
state their finding is consistent with findings based on long-term monitoring data, that
DOC increases in northern European and North American surface waters are
substantially attributable to regional decreases in acidifying, primarily S, deposition
(Montcith et al.. 2007; Evans et al.. 2005). New studies also report that N addition has
caused inconsistent changes to DOC, with N addition causing both increases (Driscoll et
al.. 2016; Liu and Greaver. 2010) and decreases (Fuss et al.. 2015; Hagedorn et al..
2012).
In the Adirondack Long-term Monitoring Program, Driscoll et al. (2016) measured
increasing DOC in 29 of 48 monitored lakes as lake pH increased. In contrast, increases
in DOC concentrations in a soil solution and streamwater studies were observed with
recovery from acidifciation in the Hubbard Brook Experimental Forest, NH (Fuss et al..
2015). Fuss et al. (2015) reviewed European and U.S. research literature, found studies
with similar observed trends, and identified a number of potential factors leading to this
difference. While surface water DOC generally emanates from soil water DOC (with the
possible exception of snowmelt), soil ionic strength and soil Al pools, as well as soil
depth, may also influence DOC in soil solution (Lofgren et al.. 2010). Decreasing ionic
strength can decrease DOC; higher levels of organic aluminum complexes in soil can
increase DOC solubility; and DOC in forest floor soil water may decrease at a faster rate
than in mineral soil water. The authors did note that Hubbard Brook's rate of decreasing
DOC concentrations appeared to be diminishing in the later 15 years of reporting.
Although most studies are on forests, Sawicka et al. (2017) used the MADOC model for
three forested and three grassland and heath sites and observed that the acidifying effect
of S deposition was the predominant control on the observed soil water DOC trends
4-65
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
(Sawicka et al.. 2017). The relative importance of S and N loading depended on soil
sensitivity to acidification and on N limitation. In all N limited soils investigated, the
modeled DOC increases over the monitoring period were dominated by the effects of
recovery from acidification (higher DOC solubility), but the effects of N enrichment
driving higher DOC production may have been important in the longer term. In contrast,
reductions in nonmarine chloride deposition and the effects of long-term warming
appeared to have been relatively unimportant.
Other recent work has further explored the mechanisms for how N addition (including
nutrient as well as acidification effects) affects DOC (Freedman and Zak. 2014; Kopacek
et al.. 2013). Freedman and Zak (2014) reported that atmospheric N deposition led to less
microbial biodiversity and favored bacterial species in the forest floor that caused lead to
reduced decomposition, increased soil C storage, and accelerated phenolic DOC
production (Figure 4-8). Kopacek et al. (2013) observed that N addition, together with
S042 deposition, increases the availability of electron acceptors for soil microbial
processes. The chemical and microbial responses include an increase in bacterial
mineralization creating DOC, increased N mineralization where N availability increases
(with an increase in bacterial uptake of DOC), chemical suppression of DOC solubility
and bioavailability in the progressively acidifying soils, and finally, the C limitation of
microbial metabolism. Bacterial assimilation of NO3 , which depends on the
stoichiometric NO;, to DOC ratio in the substrate, may initially increase with increasing
NO;, concentrations, but then decrease due to a lower pool of bioavailable DOC.
Atmospheric N Deposition
Bacterial Melaboilsm of Lignln and Humics
LMCO-nartarlng Bacteria & Acboazaciena
I*
\®
Decomposition
Ratfrs
Organic Matter
Accumulation
atler
t ion
N = nitrogen; LMCO = Laccase-like multicopper oxidase.
Source: Freedman and Zak (2014).
June 2018
4-66
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Figure 4-8 Conceptual diagram of positive (solid arrows) and negative
(dashed arrows) fluxes in nitrogen pools (squares) and carbon
pools (ovals) and the biological processes (no border) that are
affected by experimental nitrogen deposition.
4.3.10. Belowground Carbon Pools
In general, N enrichment influences C flux, C partitioning, and the amount of C
sequestered by ecosystems. A number of studies have suggested that C sequestration
increases with increasing N supply based on the changes in aboveground net primary
production [ANPP; (Le Bauer and Treseder. 2008; Xia and Wan. 2008)1. However, about
half of the C fixed annually by terrestrial vegetation is allocated to belowground pools.
Many studies have shown that the belowground C cycle does not always mirror the
aboveground cycle. For example, elevated CO2 and N both have been shown to increase
aboveground biomass production (LeBauer and Treseder. 2008). However, increases in
aboveground plant production and greater aboveground litter inputs do not necessarily
increase mineral soil C storage (Talhclm et al.. 2009; Gielen et al.. 2005; Lichtcr ct al..
2005). and increases in soil C do not necessarily result from greater aboveground litter
inputs (Prcgitzcr et al.. 2008). This disparity indicates that it is inappropriate to
extrapolate from aboveground responses to belowground processes. Aboveground plant
biomass, once dropped from the canopy, is one of the major contributors to soil organic
matter accumulation (Sullivan et al.. 2007a). ANPP generally increases with increasing N
supply (LeBauer and Treseder. 2008; Xia and Wan. 2008). However, this increase in C
input may not cause an increase in belowground ecosystem C storage. This is because the
output of C from the soil may be stimulated by N induced alteration of plant tissue
chemistry or the ratio of root: shoot, each of which will change the rate of decomposition
(see Appendix 4.3.7). Since the 2008 ISA was published there have been several new
meta-analyses on the effects of N addition on belowground carbon pools (Jansscns et al..
2010; Liu and Greaver. 2010). each integrating information from numerous pools to
improve the understanding of how N addition alters the carbon cycle belowground.
Additional studies are summarized in Table 4-11.
Liu and Greaver (2010) synthesized data from multiple terrestrial ecosystems to quantify
the response of belowground C cycling under N addition (Figure 4-9). N addition
increased aboveground litter inputs (+20%), but fine roots litter inputs were unchanged.
N addition inhibited microbial activity, as indicated by a reduction in microbial
respiration (-8%) and microbial biomass carbon (-20%). Although soil respiration was
June 2018
4-67
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
not altered by N addition, dissolved organic carbon concentration increased (+18%),
suggesting C leaching loss may increase. N addition increased the C content of the soil
organic horizon (+17%) but not the mineral soil. The increase in organic horizon C was
attributed to both increased litter input and decreased decomposition [inferred from the
lower microbial respiration rates (Liu and Greaver. 2010)1.
Field studies provide additional support for the finding that soil carbon responds
differently to N addition depending on the organic (vs. mineral) content of the soil, with
organic content tending to decrease with increasing soil depth. In four northern hardwood
forests spread across Michigan that received experimental N deposition (additional
30 kg N/ha/yr) for 10 years, Pregitzer et al. (2008) and Zak et al. (2008) reported
significant increases in soil C, particularly within the soil organic horizon. These
increases occurred despite no increase in aboveground litter production or root turnover.
A ' "NOs -labeling experiment revealed that N accumulated in SOM by first flowing
through soil microorganisms to plants, which then shed the leaves back to the detrital
layer. In a meta-analysis including data from 72 north temperate forests, (Nave etal..
2009a) found N addition increased soil C (+7.7%). The forest floor responded differently
than the deeper mineral soil layer (5- to 100-cm depth) as increased soil C storage
occurred only in the mineral soil (+12.2%), in contrast to the findings of Liu and Greaver
(2010) and Pregitzer et al. (2008).
4-68
DRAFT: Do Not Cite or Quote

-------
ANPP = aboveground net primary production; DOC = dissolved organic carbon; ER = ecosystem respiration; GPP = gross primary
productivity; MBC = microbial biomass carbon; n.s. = nonsignificant; Raboveground = aboveground respiration; RautotroPhic = soil
autotrophic respiration; Rmio-obiai = microbial respiration; Rsoii = total soil respiration,
f data from LeBauer and Treseder (2008); * data from Xia and Wan (2008); * data from Treseder (2004).
Source: Liu and Greaver (2010).
Figure 4-9 Estimation of the changes in carbon budget of terrestrial
ecosystem under nitrogen addition.
1
2	A new study has evaluated the consistency of N cycling in different ecosystems by
3	synthesizing data across five biomes in California. The results indicated that, across
4	biomes, N concentration in soil has a strong positive correlation to SOC (Yang et al..
5	2017). There are also new studies indicating N deposition increases SOM accumulation
6	without altering the biochemical composition (Zak et al.. 2017). New studies on how the
7	response of belowground soil C to N is modified by temperature and precipitation has
8	been published since the 2008 ISA (Ni et al.. 2017; Greaver et al.. 2016) and is
9	summarized in Appendix 4.7.
June 2018
4-69
DRAFT: Do Not Cite or Quote

-------
Table 4-11 Belowground carbon pools.
Process
Type of Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
soc
Soil
Great Britain
0-2 g N/m2/yr None
Model: Using model N14CP for
data from -2,000 N limited field
sites, it was predicted that, N
deposition increased NPP between
the Years 1750 and 2010,
increasing via detritus SOC by
1.2 kg C/m2 (-10%). (The authors
assumed -30% error of estimated
values.)
Tipping et al.
(2017)
SOC
Five biome classes in
California: desert,
grassland, shrubland,
forest, wetland
California
0.6-18.4 kg
N/ha/yr
Grassland soil:
76.4 kg N/ha/yr as
NPK (29:3:4);
desert soil: 60 kg
N/ha/yr as NH4NO3;
shrubland soil:
60 kg N/ha/yr as
NH4NO3
Synthesis: Across biomes, total N Yang et al. (2017)
concentration was strongly
correlated to SOC (R2= 0.88;
log(y) = [0.81 x log(x)]-1.10).
SOC
Hardwood forest birch
(Betula platyphylla) and
aspen (Populus
davidiana)
Jilin province,
China
23 kg N/ha/yr
0 kg N/ha/yr, 25 kg
N/ha/yr, and 50 kg
N/ha/yr
N Addition: This was a 6-yr
fertilization study. The authors
observed that fertilization
decreased the fraction of
macroaggregrates (2-8 mm) and
increased the fraction of
0.053-2 mm aggregates.
(P = 0.01). They concluded that N
deposition (as simulated with
fertilization) can increase formation
of micro- and macroaggregates
within macroaggregate soil and
thus stabilize C.
Chen et al. (2017)
June 2018
4-70
DRAFT: Do Not Cite or Quote

-------
Table 4-11 (Continued): Belowground carbon pools.
Process
Type of Ecosystem
Region
Ambient N/S
Deposition
kg/ha/yr
N/S Addition
kg/ha/yr
Effect of Deposition
Reference
HERO ID
Soil organic matter
Sugar maple (Acer
saccharum)-dom\riated
northern hardwood
forests
North-south
geographic range
of sugar
maple-dominated
upper Great
Lakes and
eastern North
America
6.8-11.8 kg
N/ha/yr
(wet + dry)
across sites
30 kg N03"-N/ha/yr
(NaNOs)
N Addition: SOM accumulated as
soil particulate organic matter
occlusion at N addition sites.
Zaketal. (2017)
Abundance of
carboxyl, aryl,
O/N-alkyl, and alkyl
C in mineral soil
(biochemical
composition of
forest floor
[Oe/Oi-horizon
~4 cm] and organic
matter)
Sugar maple (Acer
saccharum. )-dominated
northern hardwood
forests
North-south
geographic range
of sugar
maple-dominated
upper Great
Lakes and
eastern North
America
6.8-11.8 kg
N/ha/yr
(wet + dry)
across sites
30 kg N03"-N/ha/yr
(NaNOs)
N Addition: Abundance of
carboxyl and aryl C in forest floor
differed across sites but the
abundance of O/N-alkyl and alkyl C
did not (P= 0.10-0.80). No
difference was observed in the
proportion of aryl, O/N-alkyl, or
alkyl C in mineral soil between
ambient deposition and N addition
sites. (Particulate organic matter's
increased occlusion of organics
may shift fungal and bacterial
activity toward partial oxidation of
organics without changing
biochemistry of organic matter.)
Zaketal. (2017)
June 2018
4-71
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
4.3.11. New Biogeochemical Indicators
In this section, we report on new biogeochemical indicators reported in the literature
since the 2008 ISA. Recent literature offering potential new indicators of biogeochemical
change due to N and/or S deposition included fungal-to-bacterial ratio (Hogbcrg et al..
2013) and syntaxonomic associations rfWamelink et al.. 2011) Table 4-121.
Hogberg et al. (2013) proposed soil microbial community indices as predictors of soil
solution chemistry and N leaching in Picea abies spruce forests in southern Sweden.
Stands with low concentrations of NO3 and Al3+ had higher fungi:bacteria ratios
compared with stands with higher concentrations of NOs and Al. They identified three
promising microbial community indices as indicators of N leaching from forests; the soil
fungi:bacteria ratio was the most important.
In a study in the Netherlands, Wamelink et al. (2011) examined whether the abiotic
ranges of syntaxonomic units (associations) in terms of pH and NOs concentration can
be estimated and then, in principle, used to estimate critical loads for acid and N
deposition. They used splines to estimate abiotic ranges of syntaxonomic units on the
basis of measured soil pH and NO3 concentration and vegetation releves. They
acknowledge it is not yet possible to directly estimate ranges for syntaxa for pH and NO3
on a large scale using this approach; however, indirectly estimated soil pH and NO3
concentrations are sufficiently available to derive ranges for many associations.
June 2018
4-72
DRAFT: Do Not Cite or Quote

-------
Table 4-12
New biogeochemistry indicators.



Endpoint
Nutrient
Enrichment Acidification
Indicator Indicator
Deposition
kg/ha/yr
Addition
kg/ha/yr
Description/Direct Effect of
Soil/Water Endpoint On Biological
Effect
Reference
HERO ID
N leaching
X
Throughfall N
includes wet and
dry inputs and
ranged from 2.7 to
19
20 (NH4NO3)
Microbial community composition in
the mor layer of spruce forests and
soil solution chemistry below the
rooting zone was highly correlated.
Stands with low concentrations of
NO3" and Al were fungi dominated
and had a higher fungi:bacteria ratio
compared with stands with high
concentrations of these leachates.
Leaching stands had higher
abundance of Gram-positive
bacteria.
Hoabera et al.
(2013)
Acknowledge the approach is not yet Wamelink et al.
possible to directly estimate ranges (2011)
for syntaxa for pH and NO3" on a
large scale; however, indirectly
estimated soil pH and NO3"
concentrations are sufficiently
available to derive ranges for many
associations.
Al = aluminum; ha = hectare; kg = kilogram; N = nitrogen; NH4NO3 = ammonium nitrate; N03 = nitrate; yr = year.
Vegetation	X	X	Not specified None
health
June 2018
4-73
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
4.3.12. Differential Effects of Reduced and Oxidized nitrogen (N)
Some biogeochemical process responses to N deposition have the potential to vary
depending on whether the dominant form of deposited N was oxidized or reduced. We
have focused on presenting meta-analyses as they synthesize the central tendencies of
many studies. In some cases, the results for a given endpoint are not consistent among
meta-analyses. Table 4-13 summarize the studies that look at the most common forms of
N.
Additional studies include the observation by Evans et al. (2008) that DOC
concentrations responded predictably to the chemical form of N used for manipulation,
increasing with NaNC>3 additions or gaseous NH3+ exposure, and decreasing with most
NH4+ salt additions (Appendix 4.3.9). The authors cite the effect of the chemical form N
on acidity as a plausible mechanism. Ramirez et al. (2010a) investigated whether soil
microbes in three distinct soils (from aspen, pine, and grassland ecosystems) respond
differently six different reduced and oxidized N species, including NH4NO3, (NTh^CO
(urea), KNO3, NH4CI, (NH^SO^ and Ca(NC>3)2 (also discussed in Appendix 6). The
authors found that all inorganic forms of N fertilizer significantly decreased microbial
CO2 production but that organic N (urea) decreased the respiration rate in forest soil only
by 27% (aspen) and 11% (pine) and increased the grassland microbial respiration rate by
20%. They added that the soil pH change resulting from N addition did not appear to
influence the observed decrease in the soils' microbial respiration rate.
Jovan et al. (2012) monitored tree trunk pH during their research on eutrophic lichen
abundance in the Los Angeles air basin, oxidized forms of N (particularly HNO3)
dominate dry deposition. They noted that NH3 deposition would normally raise bark pH
but in the arid, hot Mediterranean climate of the study area, high trunk pH dust raises
bark pH.
June 2018
4-74
DRAFT: Do Not Cite or Quote

-------
Table 4-13 The effects of different forms of inorganic nitrogen on biogeochemical processes and indicators
according to meta-analyses. See Table 6-1 for the effects of different forms of inorganic nitrogen on
biological endpoints.
Process/Indicator
NO3-
NH4NO3
NH4+
Urea
UAN
Reference
Total soil C
NS
NS
NS
NS
NS
Liu and Greaver (2010)
Total soil C
NS
T
NS
NS
-
Yueetal. (2016)
Soil DOC
T
T
NS
NS
-
Yueetal. (2016)
Soil DOC
T
NS
NS
T
-
Liu and Greaver (2010)
Organic layer C
NS
T
-
T
-
Liu and Greaver (2010)
Soil respiration
NS
NS
NS
NS
NS
Yueetal. (2016)
Soil respiration3
NS
NS
T
NS
-
Liu and Greaver (2010)
CH4 emission
NS
NS
NS
NS
NS
Liu and Greaver (2009)
CH4 uptake




NS
Liu and Greaver (2009)
N2O
T
T
T
T
T
Liu and Greaver (2009)
N Recovery
80%
85%
53%
-
-
Templer et al. (2012)
aNot significant effect of the N mean for this endpoint; however, there was a significant difference among N forms.
June 2018
4-75
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
4.4.
Soil Monitoring and Databases
There are several new studies using long-term monitoring data sets in the U.S. and
Europe (Table 4-14). New studies in the U.S. include an analysis of 45 years of
biogeochemical monitoring data (Yanai et al.. 2013) and sulfur accumulation (Mitchell
and Likens. 2011) at the Hubbard Brook Experimental Forest LTER, NH. At the Niwot
Ridge LTER site, CO (Lieb et al.. 2011). identify the effects of a decade of simulated N
deposition in the southern Rocky Mountains.
Yanai et al. (2013) evaluated 45 years of biogeochemical monitoring data at the Hubbard
Brook Experimental Forest LTER site, New Hampshire. Since 1992, the ecosystem
shifted to a net N sink ~8 kg N/ha/yr. There are several possible explanations: (1) gaseous
N fluxes from the ecosystem in response to denitrification, (2) a budget discrepancy in
the net error of the other measured and estimated stocks and fluxes, and/or (3) N
accumulation in an unidentified ecosystem compartment.
As discussed in Appendix 4.3.3. Mitchell and Likens (2011) examined the sulfur
accumulation observed in over four decades of continuous long-term records for four
watersheds in Hubbard Brook Experimental Forest, NH and found that as S deposition
declined, soil moisture became a more powerful control on S release from soils than did
deposition.
At the southern Rocky Mountains Niwot Ridge LTER site, CO, U.S., Lieb et al. (2011)
found a decade of simulated N deposition to alpine ecosystems caused ongoing changes
in diversity and soil biogeochemistry, including lower soil acid buffering capacity,
decreased concentrations of exchangeable Mg2+, and increased concentrations of the
potentially toxic cations Mn2+ and Al3+. Their results suggested an N deposition threshold
for the onset of acidification at this site of 28 kg N/ha/yr.
June 2018
4-76
DRAFT: Do Not Cite or Quote

-------
Table 4-14
Biogeochemistry monitoring and databases.


Process/


Deposition

Reference
Indicator
Type of Ecosystem
Region
(kg/ha/yr)
Addition (kg/ha/yr) Effects
HERO ID
Base cation (Be)
Alpine soils
Niwot Ridge in the
6 to 8 (10 yr)
8, 28, 48, and 68 Addition: Changes in diversity,
Liebet al. (2011)
release

southern Rocky

(NH4NO3) lower soil acid buffering capacity,

soil [Al]

Mountains, CO,

decreased concentrations of Mg2+,

soil [Mn]

U.S.

and increased concentrations of

soil pH



the potentially toxic cations Mn2+





and Al3+. Results suggested an N





deposition threshold of





28 kg N/ha/yr.

Monitoring: Over four decades of Mitchell and
data were used to evaluate S Likens (2011)
budgets. Current declining inputs
of atmospheric S and the higher
outputs of SO42" in drainage
waters relative to precipitation
inputs are driven by the S stored
in the soil. Climatic change will
potentially increase SO42"
mobilization and hence could slow
the resultant recovery from
acidification.
Monitoring: Since 1992, the Yanai et al. (2013)
ecosystem shifted to a net N sink,
either storing or denitrifying
~8 kg N/ha/yr.
Sulfate leaching	Northeastern forest HBEF, NH, U.S. Not specified None
[S0421
N accumulation C:N Northern hardwood HBEF, NH, U.S. ~7	None
soil [N]	forest	(since 1992)
June 2018
4-77
DRAFT: Do Not Cite or Quote

-------
Table 4-14 (Continued): Biogeochemistry monitoring and databases.
Process/
Indicator
Type of Ecosystem
Region
Deposition
(kg/ha/yr)
Addition (kg/ha/yr)
Effects
Reference
HERO ID
NO3" leaching C:N
(lake)
C:N (mineral soil layer)
TOC:TN
Boreal soils and
lakes
Sweden
<3 to 17
None
Gradient: Significant relation
found between C:N ratios ofthe
organic soil layer and the ones of
lake waters. Evidence found of N
deposition having depressed the
C:N ratios of lake waters more
than the ones of organic soil
layers. Clear sudden increase
seen in NO3" leaching in regions
where N deposition exceeded
7.5 kg/ha/yr.
Khalili etal. (2010)
Oxidation of organic S
in humic soils due to S
dep
Ratio of reduced to
oxidized organic S
Grassland
Rothamsted Park
Grass Experiment,
Herfordshire,
England
Not specified
None
Monitoring: Analysis ofthe
effects of atmospheric SO2
emissions since the late 1800s
found acidification led to a
depletion of exchangeable Ca and
Mg and an 8x increase in
exchangeable Al.
Lehmann et al.
(2008)
Be, N, and SO42" flux
soil solution [Al]
soil solution Be
soil solution [Inorganic
N]
soil solution [SO42 ]
molar Bc:AI in soil
solution
Forest, including
beech and Norway
spruce
Switzerland, Jura
Mountains in
southern Alps. The
forests were not
managed during
the whole
observation period
Between 2000 and None
2007, mean
(-0.02 to 1.99
kmol/ha/yr)
Monitoring: A decade
(1995-2007) of monitoring data
indicates acidifying deposition
significantly decreased at three
out ofthe nine study sites due to a
decrease in total N deposition. In
the soil solution, no trend in
concentrations and fluxes of Be,
SO42", and inorganic N were
found at most soil depths at five
out ofthe seven sites, suggesting
that the soil solution reacted very
little to the changes in
atmospheric deposition.
Pannatier et al.
(2011)
June 2018
4-78
DRAFT: Do Not Cite or Quote

-------
Table 4-14 (Continued): Biogeochemistry monitoring and databases.
Process/	Deposition	Reference
Indicator	Type of Ecosystem	Region	(kg/ha/yr)	Addition (kg/ha/yr)	Effects	HERO ID
pH, ANC, base cations, Forest and streams Hubbard Brook Variable	None	Field Observation: Soil solution Fuss et al. (2015)
Ali, Alo	Experimental	monitoring 1984-2011 at HBEF
Forest, NH	showed that: pH and ANC did not
change significantly in the
Oa-horizon but ANC increased in
Bs-horizon. Total base cations
decreased in Bs-horizon. Al0
decreased and. Ali decreased in
the 2 high elevation
subwatersheds.
Al = aluminum; Ah = inorganic aluminum; Al0 = organic aluminum; Be = base cation; C = carbon; Ca = calcium; ha = hectare; HBEF = Hubbard Brook Experimental Forest;
kg = kilogram; kmol = kilomole; Mg = magnesium; Mn = manganese; N = nitrogen; NH4N03 = ammonium nitrate; N03" = nitrate; S = sulfur; S02 = sulfur dioxide; S042" = sulfate;
TN = total nitrogen; TOC = total organic carbon; yr = year.
June 2018
4-79
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
There are new studies evaluating long-term monitoring in Europe, including those
looking at S dynamics in England (Lchmann et al.. 2008) and N and S dynamics in
Switzerland (Pannaticr et al.. 2011). In Sweden, Khalili et al. (2010) examined N and C
interactions between boreal soils and lakes. Lehmann et al. (2008) found that soil
acidification from the input of oxidized organic S proceeded much more rapidly than did
recovery the reversal after reductions of atmospheric emissions and deposition. Pannatier
et al. (2011) examined monitoring data at Swiss Long-Term Forest Ecosystem Research
sites. The results suggested that the fluxes of Be, SO42 , and inorganic N in soil solution
reacted very little to the changes in atmospheric deposition. A stronger reduction in base
cations compared to Al3+ was detected at two sites, possibly indicating that acidification
of the soil solution was proceeding faster at these sites than the other sites. In Sweden,
Khalili et al. (2010) examined samples collected in selected years between 1993 and
2005. Although they found a significant relation between C:N ratios of the organic soil
layer and lake waters, the large-scale variations in soil C content were not directly linked
to C concentrations in lake waters. Soil N seems to have leached in small amounts from
the soils directly into the lakes in the form of NO;, . NO3 leaching showed a clear and
sudden increase in regions where N deposition exceeded 7.5 kg/ha/yr.
4.5. Models
The 2008 ISA described the most commonly used soil biogeochemistry models in the
U.S. used to track N and/or S deposition (Section A.3 of the 2008 ISA). The focus of
Appendix 4o is to update available information on several key models currently being
used in the U.S. to assess the effects of S and N deposition on terrestrial ecosystem soil
biogeochemistry. Steady-state models include steady-state simple mass-balance (SMB).
Dynamic models include the Very Simple Dynamic (VSD) soil acidification model,
MAGIC, PnET/BGC, and DayCent-Chem. One important input to these models is
estimating base cation weathering (BCw), and there are new updates on two methods to
estimate this parameter: Soil Texture Approximations (STA) and PROFILE.
4.5.1. Updates to Key Previously Identified Models
4.5.1.1. Soil Texture Approximation (STA) and PROFILE: Estimating Base
Cation Weathering
Base cation weathering (BCw) rate is one of the most influential yet difficult to estimate
parameters in the calculation of critical acid loads of N and S deposition for terrestrial
June 2018
4-80
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
systems. As discussed in Appendix 4.3. acidifying deposition causes base cation leaching
from soil. The supply of base cations is largely replenished by base cation weathering of
minerals from rock within the ecosystems. Obtaining accurate estimates of weathering
rates is difficult because weathering is a process that occurs over very long periods of
time. There are some new studies on estimating BCw, including a study on the clay
correlation-substrate method and PROFILE (Koseva et al.. 2010). a new application of
PROFILE in the U.S. (Phclan et al.. 2014). and an evaluation of uncertainty in estimating
BCw (Futtcr et al.. 2012).
The clay correlation-substrate method, also called clay-based Soil Texture
Approximation (STA), is an empirical steady-state model that has been used to estimate
BCw rates for forest ecosystems in the U.S. because it is simple and has low data
requirements. The STA method has an empirical function that was first developed for
European soils and later adapted to soils in Canada and the U.S. The method estimates
total base cation weathering rates (BCw, sum of Ca2+, Mg2+, Na+, and K+), based on the
relationship between established weathering classes and clay content of soils for different
acidity classes—acidic, intermediate, and basic. An alternate steady-state model,
PROFILE (Sverdrup and Warfvinge. 1993). may offer an improved method to estimate
BCw rates. It is a transferable, process-based model that simulates the weathering rates of
groups of minerals. PROFILE was developed in Sweden and is a mechanistic,
steady-state kinetics model that calculates the weathering of Ca2+, Mg2+, K+, and Na+ in
each horizon within a soil profile based on mineral specific chemical dissolution rates
and site and soil conditions. Because PROFILE is a steady-state model, weathering rates
calculated by PROFILE can be used with the SMB critical acid load model. One of the
main limitations that has discouraged the wider adoption of PROFILE for BCw rate
determinations in the U.S. is the large data requirements of the model (Koscva et al..
2010).
Koseva et al. (2010) reported on their evaluation and revision of the STA model for use
in Canadian forests. The authors compared the performance of the STA model to
PROFILE weathering estimates for soils at 75 sites in Canada. The relative ability of the
STA and PROFILE models to provide reasonable weathering rates was evaluated using
base cation mass balances at a subset of sites in Ontario (n = 19). Mineral weathering
rates calculated with the STA method at 75 sites in Canada (6-367 eq/ha/yr) were up to
38 times lower (7 times on average) than rates estimated with PROFILE
(143-2,119 eq/ha/yr,/? < 0.0001). Despite deviations from the 1:1 line, weathering
estimates obtained using the STA method were significantly correlated using PROFILE
weathering estimates (p < 0.05). The authors concluded that the "revised" STA model
they used may be more widely applicable in Canada, but not necessarily suited to all
regions in the U.S. The uncertainty of the model is largely unknown, and the three
4-81
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
equations of the clay correlation-substrate method require recalibration or revision when
transferred to new locations. In addition, the STA equations were derived for young soils
that developed following the Late Wisconsin glaciation (Koscva et al.. 2010); the clay
correlation-substrate model, which is based on clay content and parent material acidity,
may not be suitable for older, more weathered soils that were not affected by the most
recent glaciation and which cover the majority of the U.S. (U.S. EPA. 2009c).
PROFILE requires over 26 time series or site parameters as model inputs. The U.S.
Geological Survey (USGS) recently completed a soil geochemical and mineralogical
survey of the U.S. as part of the North American Soil Geochemical Landscapes Project.
This project, hereafter referred to as the USGS Landscapes Project, included mineralogy
analyses conducted on soil samples collected from 4,871 evenly spaced sites across the
U.S. (Smith et al.. 2013). These new data allow the PROFILE model to be applied at a
larger scale. Phelan et al. (2014) evaluated PROFILE using national data sets as a method
to estimate BCw rates for forests in the U.S., focusing on Pennsylvania as the first test
state. The model was successfully applied at 51 forested sites across Pennsylvania.
Weathering rates ranged from 11.9 to 924.5 meq/m2/yr and were consistent with soil
properties and regional geology. The authors suggest that the method be applied to other
locations to further evaluate the performance of the model.
A data set of published BCw rates was evaluated by Futter et al. (2012). The data set
included 394 individual silicate mineral weathering rate estimates from 82 poorly
buffered, silicate-mineral-dominated locations across the world where at least 3 published
estimates of BCw were available. The researchers found uncertainty for the input data
high relative to the estimated contribution of model parameter uncertainty of BCw
weathering rate (meq/m2/yr) to overall variability related to PROFILE and MAGIC,
ranging from 1.7 to 2.1 for the two models, respectively. The two models have been most
widely used to make assessments of weathering rates for environmental decision making,
and each represents the landscape differently. PROFILE provides estimates for shallow
one-dimensional soil, whereas MAGIC integrates weathering processes for the whole
catchment, including deep soil weathering. Therefore, lower weathering rate estimates
are shown from PROFILE in comparison to MAGIC, which considers deep soil
weathering to constitute a potentially important source of Be in catchment weathering
(Futter etal.. 2012).
Whitfield (2018) evaluated major sources of uncertainty associated with using PROFILE
for upland forests in the continental U.S. Mineral stoichiometry was not an important
influence on BCw estimates (uncertainty <1%). Characterizing B-horizon mineralogy by
averaging A- and C-horizons was found to be a minor (< 5%) contributor to uncertainty
in some areas, but where mineralogy is known to vary with depth the uncertainty can be
4-82
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
large. Estimating mineral-specific surface areas had a strong influence on estimated
BCw, however the greatest uncertainty in BCw estimates, was due to the particle size
class-based method used to estimate the total specific surface area upon which
weathering reactions can take place.
4.5.1.2. Steady-State Mass Balance
The 2008 ISA documented a model to assess CLs for acidification in forest soils based on
simple mass-balance equations [SMBE; (McNultv et al.. 2007)1. This study estimated
critical acid load and exceedance in soils at a 1-km2 spatial resolution across the U.S. A
second publication discussed the uncertainties associated with this model and
national-scale assessment (Li and McNultv. 2007). The authors quantified uncertainty
under natural variability in 17 model parameters and determined the relative
contributions of each in predicting critical loads. The results indicated that uncertainty in
the CLs came primarily from components of base cation weathering (BCw; 49%) and
acid neutralizing capacity (46%), whereas the most critical parameters were BCw base
rate (62%), soil depth (20%), and soil temperature (11%). The authors concluded that
improvements in estimates of these factors are crucial to reducing uncertainty and
successfully scaling up SMBE for national assessments. This work remains the best
national-scale estimate of terrestrial soil acidification in the U.S. (see Appendix 4.6). A
new study by Posch etal. (2011) is a regional application of SMB models, while several
other studies using SMB have been published to determine acidification critical loads
based on critical limits of Bc:Al and ANC (Phelan et al.. 2014; Duarte et al.. 2013; Jung
et al.. 2013; Whitfield and Watmough. 2012; Forsius et al.. 2010; McNultv and Boggs.
2010; Nasr etal.. 2010). The biological implications of these critical loads are discussed
in Appendix 5 and a summary of SMB CLs in the U.S. is presented in Appendix 4.6.
Posch et al. (2011) published an article outlining SMB models for use in N deposition on
ecosystem biodiversity. This approach has the well-established benefit of easy regional
applicability, while incorporating specified critical chemical criteria to protect specified
receptors. Rather than indicating an upper limit for deposition (i.e., critical load), linked
nutrient nitrogen and acidity chemical criteria for plant occurrence produce an "optimal"
nitrogen and sulphur deposition envelope. This method is similar to the methods
developed in the 2012 Policy Assessment for the Review of the secondary NAAQS for
Oxides of N and S.
June 2018
4-83
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
4.5.1.3.
ForSAFE and ForSAFE-VEG
The ForSAFE model (Wallman et al.. 2005) is the biogeochemical simulator platform
that simulates the cycles of carbon, nitrogen, base cations (Be), and water in a forest
ecosystem, including simulation of soil chemistry, tree growth, and soil organic matter
accumulation or depletion. ForSAFE requires site-specific inputs of the physical
properties of the soil (including mineralogy, hydrological parameters, density, depth, and
stratification), tree type, and time series of atmospheric deposition and climatic data
(temperature, light, and precipitation). The model gives monthly estimates of weathering
rates, soil moisture, soil solution concentrations, uptake fluxes of N and Be, litterfall,
decomposition and mineralization, as well as photosynthesis and growth rates.
ForSAFE-VEG is a composite model, in which the VEG module (Sverdrup et al.. 2007)
reads a set of five drivers (soil solution pH, Be concentration, N concentration, ground
level light, soil moisture) from ForSAFE, including air temperature, and uses them to
estimate the relative abundance of a set of indicator plants at the site. The result is a
model chain that can link changes in atmospheric deposition, climatic conditions, and
land use to responses in the biogeochemistry and plant community composition at the site
level, both historically and in the future. A new study by Belvazid et al. (201 la) revealed
limitations in the model simulation of N concentrations in soil solution. The authors
concluded that the biogeochemical model platform must be improved to simulate N
processes more accurately before it is used to calculate CL for N deposition. The model
overestimated the actual N concentrations in the soil solution. Several new applications
of ForSAFE-VEG have been published (Sverdrup et al.. 2012; Belvazid et al.. 2011a).
and the results are discussed in Appendix 5 and Appendix 6.
4.5.1.4. Model of Acidification of Groundwater in Catchment (MAGIC)
The MAGIC (Model of Acidification of Groundwater in Catchment) (Cosby et al..
1985a; Cosby et al.. 1985c; Cosby etal.. 1985b) is one of the most well-known dynamic
models of aquatic and terrestrial acidification. It is a lumped-parameter model of soil and
surface water acidification in response to atmospheric deposition based on process-level
information about acidification. "Lumped-parameter" refers to the extent that spatially
distributed physical and chemical processes in the catchment are averaged or lumped
together without affecting the model's reproduction of catchment response. Process-level
information refers to how the model characterizes acidification into (1) a section in which
the concentrations of major ions are assumed to be governed by simultaneous reactions
involving SOr adsorption, cation exchange, dissolution-precipitation-speciation of
aluminum, and dissolution-speciation of inorganic carbon and (2) a mass balance section
in which the flux of major ions to and from the soil is assumed to be controlled by
June 2018
4-84
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
atmospheric inputs, chemical weathering, net uptake and loss in biomass, and losses to
runoff. One strength of MAGIC is the size of the pool of exchangeable base cations in the
soil. As the fluxes to and from this pool change over time due to changes in atmospheric
deposition, the chemical equilibria between soil and soil solution shift to give changes in
surface water chemistry. The degree and rate of change of surface water acidity thus
depend both on flux factors and the inherent characteristics of the affected soils. The data
requirements to run dynamic models like MAGIC are considerable. The equations that
characterize the chemical composition of soil water in MAGIC contain 33 variables and
21 parameters. Data required to conduct dynamic modeling are not as available in as
many places as the data required to conduct steady-state modeling.
Oulehle et al. (2012) presented a new formulation of the acidification model MAGIC that
uses decomposer dynamics to link N cycling to C turnover in soils. In comparisons with
earlier versions, the new formulation more accurately simulates observed short-term
changes in NOs leaching, as well as long-term retention of N in soils. The authors state
that the new formulation gives a more realistic simulation of observed changes in N
leaching. The new formulation also provides a reasonable simulation of the long-term
changes in C and N pools (and C:N ratio) with SOM.
MAGIC has recently been used to calibrate BCw at 140 locations throughout the
southern Appalachian Mountains, where input data were sufficient for running the
dynamic model (Povak et al.. 2014). Results were then extrapolated to the region.
McDonnell et al. (2014b) used these calibrated regionalized BCw values as inputs for
regional steady-state CL modeling.
4.5.1.5. Photosynthesis and Evapo Transpiration-Biogeochemical (PnET-BGC)
The PnET-BGC (Photosynthesis and Evapo Transpiration-Biogeochemical) model is an
integrated forest-soil-water model that has been used to assess the effects of air pollution
and land disturbances on forest and aquatic ecosystems (Gbondo-Tugbawa et al.. 2001).
The model was developed by linking two submodels: PnET-CN [PnET-carbon and
nitrogen; (Aberetal.. 1997)1 and BGC (Gbondo-Tugbawa et al.. 2001). The main
processes in the model include tree photosynthesis, growth and productivity, litter
production and decay, mineralization of organic matter, immobilization of nitrogen,
nitrification (Aberetal.. 1997). vegetation and organic matter interactions of major
elements, abiotic soil processes, solution speciation, and surface water processes
(Gbondo-Tugbawa et al.. 2001). The hydrologic algorithms used in PnET-BGC are
summarized by Aber and Federer (1992) and Chen and Driscoll (2005). PnET-BGC has
the capability of using multiple soil layers (Chen and Driscoll. 2005) to model seasonal
June 2018
4-85
DRAFT: Do Not Cite or Quote

-------
1
2
variations in soil parameters and chemistry. Applications and conceptual advancements
to the model published since 2008 are summarized in Table 4-15.
4.5.1.6. DayCent-Chem
3	DayCent-Chem links two widely accepted and tested models, one of daily
4	biogeochemistry for forest, grassland, cropland, and savanna systems, DayCent (Parton et
5	al.. 1998). and the other of soil and water geochemical equilibrium, PHREEQC
6	(Parkhurst and Appelo. 1999). The linked DayCent/PHREEQC model was created to
7	capture the biogeochemical responses to atmospheric deposition and to explicitly
8	consider those biogeochemical influences on soil and surface water chemistry. The linked
9	model expands on DayCent's ability to simulate N, P, S, and C ecosystem dynamics by
10	incorporating the reactions of many other chemical species in surface water.
June 2018
4-86
DRAFT: Do Not Cite or Quote

-------
Table 4-15 Photosynthesis and Evapo Transpiration-Biogeochemical (PnET-BGC) and DayCent.
Region
Deposition
Model
(Dynamic or
Steady State)
Approach/Observation
Reference
HERO ID
Hubbard Brook
Experimental
Forest (HBEF),
NH
526 eq S/ha/yr;
327 eq N03"/ha/yr;
-522 eq
NO3- + NHs/ha/yr
PnET-BGC
Application: Combination of multiple deposition scenarios of S, NO3", and
base cation deposition (0 to 100% decrease in 20% increments after 2008), as
well as current climate and climate change scenarios to Year 2100.
Wu and Driscoll
(2010)
Hubbard Brook
Experimental
Forest, NH
Not specified but see
Wu and Driscoll
(2010)
PnET-BGC
Conceptual Advancement: PnET-BGC modified to include CO2
Pourmokhtarian et al.
(2012)
San Bernadino
Mtns., CA
Two sites:
(1)	8.8 kg N/ha/yr;
(2)	70 kg N/ha/yr
DayCent
Conceptual Advancement: DayCent modified to include O3 effects
Bvtnerowicz et al.
(2013)
Adirondack
Long-term
Monitoring
Program (ALTM;
128 lakes)
Mean wet + dry dep.
2009-2011:
S = 20-34 meq/m2/yr;
N = 24-33 meq/m2/yr
PnET-BGC
Application: Controlling S deposition is more effective in promoting acidic lake
recovery than controlling S + N deposition.
Reducing S dep. 60% beyond 2011 level is predicted to restore 28% of
impaired lakes to ANC 20 peq/L > by 2050 and 60% of lakes by 2200. An ANC
of 11 peq/L can be achieved to 53% of lakes by 2050.
Fakhraei et al. (2014)
Application: PnET-BGC was used to evaluate biophysical factors that affect Zhou et al. (2015c)
CLs and TLs of acidity for the Constable Pond watershed, as an example of a
chronically acidic drainage lake in the Adirondack region of New York, U.S.
These factors included a range of future scenarios of decreases in atmospheric
nitrate, ammonium and sulfate deposition from present to 2200.
Great Smoky 6.8-27.8 kg S/ha/yr PnET-BGC Application: Simultaneous reduction in SO42" and NO3" deposition is more Zhou et al. (2015a)
Mtns. National 6 1-16 6 kg N/ha/yr	effective at increasing stream ANC than SO42" alone. NO3" leaching continues
Park	(2004-2008)	as N deposition decreases. Stream recovery is delayed as NO3" facilitates
desorption of legacy SO42" that is adsorbed to acid soils.
Adirondack Mtns.
Region, NY
S042-= 290.3-365.7
eq/ha/yr;
N03"= 172.5-233.5
eq/ha/yr
June 2018
4-87
DRAFT: Do Not Cite or Quote

-------
Table 4-15 (Continued): Photosynthesis and Evapo Transpiration-Biogeochemical (PnET-BGC) and DayCent.
Region
Deposition
Model
(Dynamic or
Steady State)
Approach/Observation
Reference
HERO ID
Great Smoky
Mtns. National
Park
3.1 kg S/ha/yr
5.1 kg N/ha/yr
PnET-BGC
Application: Due to soil SO42" adsorption capacity, reducing NHV deposition
would be more effective in stream recovery than reducing NOx and SO2.
Fakhraei et al. (2016)
June 2018
4-88
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
4.5.2.
New Models (Published since 2008)
4.5.2.1. SMARTml
SMARTml is an assemblage of surface complexation models (SCMs), which are used in
in predicting dynamics in soil chemistry without any site-specific calibration. In the first
published application of SMARTml to a spruce forest site in Germany, model results
matched observations well overall (Bonten etal.. 2011). Simulations deviated from
observations only for soil layers or parameters for which insufficient information was
available. These positive results demonstrate the potential to further apply SCMs in
dynamic modeling. Current results only refer to a single site, and the testing of SCMs for
more cases with differences in soil types, depositions, and environmental conditions are
needed to better understand the strengths of this model.
4.5.2.2. Very Simple Dynamic (VSD) and VSD+
The VSD model (Very Simple Dynamic Soil Acidification Model) is the simplest
extension of the steady-state SMB model into a dynamic model. It does this by including
cation exchange and time-dependent N immobilization (accumulation). The VSD model
is designed for sites with little available data and for applications on a large regional or
continental scale. The model has a short execution time that allows rapid scenario
analyses and the calculation of target loads (i.e., deposition targets), which result in a
desired chemical condition in the soil (solution) in a specified year. Posch and Reinds
(2009) developed a version of the VSD for steady-state critical load applications at the
regional scale. Although simpler than other widely used dynamic models (such as
MAGIC and SAFE), VSD contains the basic physical and chemical relationships
common to all these models. However, the model's simplicity also means that some
processes have either been left out altogether (e.g., SO42 sorption) and others strongly
simplified (e.g., N cycling processes). As a consequence, VSD is not best suited for sites
where SO42 adsorption is important. Furthermore, the simple description of N processes
does not allow simulating decreasing soil N pools (and increasing C:N ratios) under
reduced N inputs.
Bonten et al. (2015) describe an extension of the VSD model, called VSD+, which
includes an explicit description of C and N turnover. The VSD+ model includes an
explicit description of organic C and N turnover, whereas in the VSD model, N
immobilization depends on the N availability and C sequestration mostly depends on the
June 2018
4-89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
N immobilization rate, which is controlled by user-prescribed C:N ratios. The VSD+
model also includes S adsorption on soils. The authors apply VSD+ to three forest stands,
which differ in N deposition and soil C:N ratios. Results suggest that at some locations
VSD+ can accurately predict trends and absolute values of NOs and NH44" concentrations
in soil and stream waters, soil C:N ratios and pH.
4.5.2.3. Soil Organic Matter
The Soil Organic Matter (SOM) model was developed as an alternative to the
decomposition module of the PnET model for application in northeastern forests. It relies
on empirical representations of litter decomposition and soil C turnover rates, and
explicitly represents multiple soil horizons. Tonitto et al. (2014) examined the effect of N
addition on SOM dynamics. Their model simulations suggested that ambient atmospheric
N deposition at the forest has led to an increase in cumulative 0-, A-, and B-horizons C
stocks of 211 g C/m2 (3.9 kg C/kg N) and 114 g C/m2 (2.1 kg C/kg N) for hardwood and
pine standards, respectively. They concluded that the model proved largely able to
simulate soil C and N dynamics at their study site under control conditions and that field
observations, mechanistic experiments, and model simulations suggest that the addition
of N to forest ecosystems could have a substantial effect on forest soil C accumulation
via suppression of organic matter decomposition.
4.5.2.4. ORCHIDEE-Carbon-Nitrogen
ORCHIDEE Carbon Nitrogen (O-CN) is a terrestrial biosphere model which has been
developed from the land surface model ORCHIDEE and describes the N and C fluxes
and stocks of vegetation and SOM for 10 natural plant functional types, as well as C3 and
C4 croplands at a half-hourly timescale. Zaehle (2013) determined that the estimate C
sequestration from the process-based O-CN model is lower than was found in earlier
studies based on simple biogeochemical models and upscaling of field-based
measurement, which have estimated C sequestration based on N deposition estimates as
0.4-0.7 Pg C/yr in 1990. The O-CN results were within the range of previous model
simulations with the current generation of C-N cycle models (0.2-0.6 Tg N/yr). The
authors believe the study provides an advance over previous assessments because it relies
on a "process-based ecosystem model that integrates the key C-N cycle interactions and
their coupling to biogeophysical processes, while considering the impacts of atmospheric
(climate, CO2) and land cover changes." Key uncertainties in the modeling include (1) the
response of canopy-level photosynthesis to N additions, (2) changes in the allocation
June 2018
4-90
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
patterns (root: shoot ratio), and (3) the competition of plants and soil microbes for the
added (or reduced) amount of N.
4.5.2.5. Dynamic Model N14C
Dynamic model N14C is a plant-soil N and C cycling model that simulates terrestrial
ecosystem responses to atmospheric N deposition (Tipping et al.. 2012). The model
includes four plant functional types: broadleaved and coniferous trees, herbs, and dwarf
shrubs. It simulates net primary production (NPP); C and N pools; leaching of DOC,
DON, and inorganic N; denitrification; and the radiocarbon contents of organic matter on
an annual time step. The model simulates annual plant growth and turnover and soil C
and N cycling and is reasonably successful at reproducing average results. However, for
individual sites, there are no significant correlations for C pools or C:N ratios and only
weak relationships for N pools and inorganic N leaching Inorganic leaching has
traditionally been considered one of the main indicators of N saturation (however, see
new studies in Appendix 4.3.2) and therefore an important goal of N14C is to simulate its
response to N enrichment.
4.5.2.6. Dynamic Simulation Model of Ecosystem Nitrogen
Perakis and Sinkhorn (2011) reported S15N constraints on long-term N balances in
temperate forests using a dynamic simulation model of ecosystem N and S15N. Their
model evaluated which combination of N input and loss pathways could produce a range
of high ecosystem N contents characteristic of forests in the Oregon Coast Range, U.S.
Ecosystem S15N displayed a curvilinear relationship with ecosystem N content and
largely reflected mineral soil, which accounted for 96-98% of total ecosystem N. Model
simulations of ecosystem N balances parameterized with field rates of N leaching
required long-term average N inputs that exceed atmospheric deposition and asymbiotic
and epiphytic N2-fixation, and that were consistent with cycles of post-fire N2-fixation by
early successional red alder. Soil water 6l5NO;, patterns suggested a shift in relative N
losses from denitrification to NOs leaching as N accumulated, and simulations identified
NO3 leaching as the primary N loss pathway that constrains maximum N accumulation.
4.5.3. Comparative Analyses
Tominaga et al. (2009) used HBEF, NH, U.S., as the setting to evaluate the performance
of three uncalibrated process-oriented models. They performed a Monte Carlo
June 2018
4-91
DRAFT: Do Not Cite or Quote

-------
1	multiple-model evaluation framework of the dynamic models MAGIC, SAFE, and VSD.
2	The greatest differences in model outputs were attributed to the cation exchange
3	submodel, with Gapon-exchange based models retaining more base cations on the
4	exchange complex and releasing less into solution, resulting in lower soil solution ANC
5	values. Given the same deposition scenario, the three models (without calibration)
6	simulate changes in soil and soil solution chemistry differently, but the basic patterns
7	were similar.
8	Bonten et al. (2015) compared common dynamic models used to evaluate soils
9	(Table 4-16). Comparisons are is summarized in Table 4-17.
June 2018
4-92
DRAFT: Do Not Cite or Quote

-------
Table 4-16 Overview of properties of four dvnamic soil chemistrv models as characterized in Bonten et al.
(2015).
Features
VSD
MAGIC
ForSAFE
SMARTml
Number of soil layers
1
1-3
1-20
1-15 soil horizons
Temporal resolution
Annual
Annual/monthly
Monthly
Variable (day-yr)
Be weathering
External input
Calibrated, or
external-input
Submodel
External input
Forest nutrient uptake
External input
External input
Submodel
Submodel
Runoff-to-precipitation ratio
External input
External input
Fixed ratio
External input
Sulfate adsorption
Not included
Langmuir isotherm
Yes
At Fe/AI hydroxides using a 2pK-DDL model
N immobilization
Fractional, fixed, or
external input
Fractional, fixed, or
external input
Submodel
Submodel
Nitrification
100%
Fractional
Submodel
Submodel
Denitrifi cation
Fractional, fixed
Fractional
None or
submodel
Submodel
Soil N build-up controlled by C:N ratio
Yes
Yes
No
Submodel
CO2 degassing in surface water
Yes
Yes
Yes
No surface water
AI(OH)3 precipitation in stream
Yes
Yes
Yes
No surface water
Lumped base cations (Ca2+, Mg2+, and K+)
Yes
No
Yes
No
DOC dissociation model
Oliver or simple
monoprotic
Triprotic
Oliver
NICA-Donnan
June 2018
4-93
DRAFT: Do Not Cite or Quote

-------
Table 4-16 (Continued): Overview of properties of four dynamic soil chemistry models as characterized in
{Bonten, 2016, 3287736@@author year}.
Features
VSD
MAGIC
ForSAFE
SMARTml
Cation exchange
Gaines-Thomas or
Gapon
Gaines-Thomas
Gapon
OM: NICA-Donnan clay: Donnan gel hydroxides: 2pk
DDL
Number of cation exchange equations
2
4
1
Not relevant
Ions in soil solution charge balance
12
28
16
Full speciation of ions in the soil solution
First appearance in the literature
2009
1985
1993
2011
June 2018
4-94
DRAFT: Do Not Cite or Quote

-------
Table 4-17
Model comparison.



Soil Process/
Indicator
Type of Ecosystem
Region
Deposition
kg/ha/yr
Modeled Observation
Reference
HERO ID
Soil S
Forest
Three Swiss forest
monitoring sites
Not specified
In a comparison of four models VSD, MAGIC,
For-SAFE, and SMARTml, ForSAFE-modeled S
concentrations tended to be underestimated by
the model.
Bonten et al. (2015)
Soil pH
Forest
Three Swiss forest
monitoring sites
Not specified
In a comparison of four models VSD, MAGIC,
For-SAFE, and SMARTml, ForSAFE results
indicated upper soil layers can have higher
modeled pH than observed.
Bonten et al. (2015)
Soil ANC and pH
Scots pine plantation
Long-term forest
monitoring site in
U.K. (Peak District
of northern
England)
Not specified
In a comparison of four models VSD, MAGIC,
For-SAFE, and SMARTml, MAGIC successfully
reproduced the trend in acidity (ANC and pH) in
soil solution over the majority of the (ca. 1990s)
monitoring period.
Bonten et al. (2015)
Soil S042"
Spruce forest
Gardsjon, Sweden
Not specified
In a comparison of four models VSD, MAGIC,
For-SAFE, and SMARTml, ForSAFE modeling
results showed SO42" was slightly
underestimated compared to measured SO42"
concentrations in soil solutions at 5-, 10-, and
20-cm depth (1990-2000).
Bonten et al. (2015)
Soil SO42"
Spruce forest
Bechtel,
Switzerland
Not specified
In a comparison of four models VSD, MAGIC,
For-SAFE, and SMARTml, ForSAFE-modeled
soil SO42" was reasonably well modeled at
20-cm depth but clearly below measured values
at 100-cm depth (1990-2005).
Bonten et al. (2015)
June 2018
4-95
DRAFT: Do Not Cite or Quote

-------
Table 4-17 (Continued): Model comparison.
Soil Process/


Deposition

Reference
Indicator
Type of Ecosystem
Region
kg/ha/yr
Modeled Observation
HERO ID
Ion
Forest
Three Swiss forest
Not specified
Comparison of four models VSD, MAGIC,
Bonten et al. (2015)
concentrations, Al

monitoring sites

For-SAFE, and SMARTml. VSD reproduced the





ion concentrations similar to measured values,





however the modeled Al concentrations were





substantially lower than measured ones. VSD





only calculates free Al3+, whereas





measurements also include other Al-species as





Al-hydroxides and Al complexed by dissolved





organic matter and fluoride.

Base cations
Spruce forest
Gardsjon, Sweden;
Not specified
ForSAFE results compared to observations. In
Bonten et al. (2015)
(nutrients

Bechtel,

Sweden, nutrient base cations were slightly

Ca + Mg + K; Na;

Switzerland; and

underestimated compared to measured values

and Ca)

long-term

in soil solution at 5-, 10-, and 20-cm depth



monitoring site in

(1990-2000), while Na+ in soil solution was



Germany

relatively well modeled at 5-, 10-, and 20-cm





depth but less well in deeper soil. In





Switzerland, base cations modeling results were





"within the range of the measured values with a





slight underestimation at 20-cm depth"





(1990-2005). In Germany, Ca2+ agreed with





measured values in soil solution (1990-2005).

Al
Spruce forest
Bechtel,
Not specified
ForSAFE-modeled Al3+ was "in the range of the
Bonten et al. (2015)


Switzerland

measured soil solution values at 20-cm depth"





(1990-2005).

Al, Ca, and
Norway spruce forest
Long-term
Not specified
SMARTml results showed that modeled Al3+,
Bonten et al. (2015)
protons

monitoring site in

Ca2+, and protons agree with measured values



Germany

observations in soil solution (1990-2005).

June 2018
4-96
DRAFT: Do Not Cite or Quote

-------
Table 4-17 (Continued): Model comparison.
Soil Process/	Deposition	Reference
Indicator Type of Ecosystem	Region	kg/ha/yr	Modeled Observation	HERO ID
Base saturation Forest	Great Smoky	5.1 kg N/ha/yr	PnET-BGC used to model 30 stream	Fakhraei et al. (2016)
Mountains National (36.5 mmolc/m2/yr) watersheds during 1981-2014 when SO42" and
Park	NO3" deposition decreased (81 and 53%, resp.).
Hindcast modeling (beginning ca. 1850) showed
decreased soil base saturation from 17.8%
(preindustrial median) to 12.6% (current
median).
AOSR = Athabasca oil sands region; S042 = sulfate; kg = kilogram; HBEF = Hubbard Brook Experimental Forest; S = sulfur; yr = year.
1
June 2018	4-97	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
4.6.
National-Scale Sensitivity
The 2008 ISA documented that by the end of the 1980s, after nearly a decade of intensive
research performed under the original National Acid Precipitation Assessment Program
(NAPAP) research program, the regions of the U.S. with acid-sensitive waters and
ecosystems were well recognized. These acid-sensitive ecosystems are generally located
in upland mountainous terrain in the eastern U.S. and are underlain by weathering-
resistant bedrock, such as granite or quartzite sandstone. The 2008 ISA documented maps
of the U.S. that showed areas sensitive to acidification. However, similar maps for areas
sensitive to the eutrophication effects of nitrogen were not available. Strong evidence was
documented demonstrating that biogeochemical sensitivity to deposition-driven
eutrophication and acidification is the result of historical loading, geologic/soil conditions
(e.g., mineral weathering and S adsorption), and nonanthropogenic sources of N and S
loading to the system. The 2008 ISA documented that there was is no single deposition
level applicable to all ecosystems in the U.S. that will describe the onset of eutrophication
or acidification. Since the 2008 ISA, there are new publications commenting on recovery
of terrestrial ecosystems at either the national scale (NAPAP. 2011) or in specific regions
(Lawrence et al.. 2015a; McDonnell et al.. 2013; Elliott et al.. 2008). One new paper
evaluates national-scale terrestrial critical loads for nitrate leaching (Pardo et al.. 201 lb),
but work on national-scale soil acidification published in 2007 remains the most recent
national assessment of this effect (McNultv et al.. 2007).
4.6.1. Acidification Recovery
It is important to note that different chemical pools within the soil may recover from
declining N and S atmospheric deposition at different rates. For example, soil solution
Ca: Al ratio or SO42 or NO3 concentration are faster responding indicators than total N.
Indicators that are linked to slow pools (such as soil percentage base saturations or soil
C-to-N ratios) will have slow response times with regard to changes in atmospheric
deposition. An indicator such as ANC which is influenced by both fast (solution) and
slow (soil) pools has an intermediate response time. In addition to how indicators convey
rates of recovery in different biogeochemical pools, recovery can be documented by
empirical evidence and projected by models of recovery trajectories.
The most recent national-scale assessment of recovery from acidifying deposition was the
National Acid Precipitation Assessment Program (NAPAP) report to Congress (NAPAP.
2011). NAPAP is a cooperative federal program first authorized in 1980 to coordinate
acid rain research and report the findings to Congress. The NAPAP member agencies are
June 2018
4-98
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
the U.S. EPA, the U.S. Department of Energy, the U.S. Department of Agriculture, the
U.S. Department of Interior, the National Aeronautics and Space Administration, and the
National Oceanic and Atmospheric Administration. The report concluded that few studies
have evaluated terrestrial ecosystem health relative to acidification effects over time, and
soils in the most acid-sensitive regions for which there is more data continue to acidify.
New evidence since the publication of the NAPAP (2011) is summarized in Table 4-18.
A new study supports the beginning of recovery from soil acidification in the
northeastern U.S. (Fuss et al.. 2015; Lawrence et al.. 2015a). Fuss et al. (2015) studied
Hubbard Brook forest in NH and found indicators of recovery of acidification in soil
solution measurements taken from 1984-2011. At the same site, Phelan et al. (2016)
conducted a modeling study that showed current reductions in deposition generally halted
further damage to soils and plants and resulted in no or only partial recovery.
In the southern Appalachian Mountains, modeling studies suggest current stress and
recovery likely to take decades even under scenarios of large reductions in S deposition
(McDonnell et al.. 2013; Elliott et al.. 2008). Rice et al. (2014) calculated for many
forested, unglaciated watersheds from Pennsylvania to Georgia will begin releasing SO42
over the next two decades (Appendix 4.3.3). Unglaciated soils, like those that occur in the
southeastern U.S., accumulate S that is slowly released from soil pools into drainage
water, a process that temporarily delays ecosystem recovery in response to decreases in S
deposition (Appendix 4.3.3).
The acid-base characteristics of DOC are an important part of understanding the recovery
potential for soils, lakes, and streams impaired by acid deposition; these effects are
discussed in Appendix 4.3.9.
4-99
DRAFT: Do Not Cite or Quote

-------
Table 4-18
Recovery.





Empirical or

Reference
Region
Type of Ecosystem
Model
Approach/Observation
HERO ID
Southern	Forest soil	NuCM	Model: Modeled three S deposition simulations: current, 50% decrease, Elliott et al. (2008)
Appalachian	and 100% increase, at Joyce Kilmer (JK), Shining Rock (SR), and Linville
Mountains, NC	Gorge (LG) wilderness areas. Low Ca:AI ratios results suggest that
forests at SR and LG are significantly stressed under current conditions.
The soil SO42" retention is low, perhaps contributing to the high degree of
acidification. The soils are very acidic and low in weatherable minerals.
Even with large reductions in SO42" and associated acid deposition, it
may take decades before these systems recover from depletion of
exchangeable Ca, Mg, and K.
Southern
Appalachian
Mountains, U.S.
65 streams and their
watersheds
MAGIC model Model: A study used the MAGIC model to evaluate soil Be status. Future
S deposition reduction scenarios (6, 58, 65, and 78% reduction), and
various changes in timber harvest, temperature, and precipitation were
modeled. Each of the scenario projections indicated that median Year
2100 soil exchangeable Ca will be at least 20% lower than preindustrial
values. The simulations suggested that substantial mass loss of soil Be
has already occurred since preindustrial times. Soil Be pools in the study
region are expected to remain significantly below preindustrial conditions
for more than 100 yr, regardless of changes in climate, S deposition, or
timber harvest.
McDonnell et al.
(2013)
Great Smoky Forest soils and	PnET-BGC	Application: Simultaneous reduction in SO42" and NO3" deposition is Zhou et al. (2015b)
Mountains	streams	more effective at increasing stream ANC than SO42" alone. NO3" leaching
National Park,	continues as N deposition decreases (this is due to unmanaged forests).
U.S.	Stream recovery is delayed as NO3" facilitates desorption of legacy SO42"
that is adsorbed to acid soils.
Great Smoky Forest soils and	PnET-BGC	Model: Modeling of 30 stream watersheds characterized by decreased Fakhraei et al. (2016)
Mountains	streams	SO42" and NO3" deposition during 1981-2014 (81 and 53%) showed that
National Park,	stream recovery has been limited and delayed due to the high sulfate
U.S.	adsorption capacity of soils in the park resulting in a long lag time for
recovery of soil chemistry.
June 2018
4-100
DRAFT: Do Not Cite or Quote

-------
Table 4-18 (Continued): Recovery.
Empirical or	Reference
Region	Type of Ecosystem	Model	Approach/Observation	HERO ID
Eastern Canada Forest and streams Recovery	Field Observation: 27 sites exposed to reductions in wet SO42"	Lawrence et al.
and the	deposition of 5.7-76%, over intervals of 8-24 yr. Results are decreased (2015a)
northeastern U.S.	exchangeable Al in the O-horizon and increases in pH in the O and
B-horizons at most sites. Among all sites, reductions in SO42" deposition
were positively correlated with ratios of base saturation and negatively
correlated with exchangeable Al ratios in the O-horizon. However, base
saturation in the B-horizon decreased at one-third of the sites, with no
increases.
Hubbard Brook
Experimental
Forest, NH
Forest and streams
Ali Recovery
Field Observation: slowed losses of base cations from soil and
decreased mobilization of dissolved inorganic aluminum were observed.
Streamwater pH at the watershed outlet increased at a rate of
0.01 units/yr and the acid neutralizing capacity (ANC) gained
0.88 [jeq/L/yr. Dissolved organic carbon generally decreased in stream
water and soil solutions. Both baseline and chronic acidification
(measured during snowment) are recovering at this site.
Fuss et al. (2015)
Hubbard Brook
Experimental
Forest, NH, and
Bear Brook
Watershed, ME
Forests soils and
plants
Model
For-SAFE-Veg
Model: Only when future deposition to 2100 was returned to preindustrial
levels was recovery of soil and plant community to 1900 conditions
projected. Policy-based reductions in deposition generally halted further
damage to soils and plants and resulted in no or only partial recovery.
Phelan et al. (2016)
Adirondack
Long-term
Monitoring
Program (ALTM;
128 lakes)
Forests soils and
lakes
PnET-BGC
Model: Controlling S deposition is more effective at acidic lake recovery
than controlling S + N deposition. Reducing N deposition is less effective
because the resulting increase in soil pH leads to soil desorption of SO42"
and other anions.
Reducing S dep. 60% beyond 2011 level is predicted to restore 28% of
Fakhraei et al. (2014)
impaired lakes to ANC 20 peq/L > by 2050 and 60% of lakes by 2200. An
ANC of 1 peq/L can be achieved to 53% of lakes by 2050.
June 2018
4-101
DRAFT: Do Not Cite or Quote

-------
Table 4-18 (Continued): Recovery.
Region
Type of Ecosystem
Empirical or
Model
Approach/Observation
Reference
HERO ID
Adirondack Mtns.
Region, NY
Forests soils and
lakes
PnET-BGC	Model: Decreasing SO42 deposition is 4.6* more effective than NO3
decrease for Years 2040-2050.
Decreases in NO3" deposition is more effective at increasing lake ANC
than an equal decrease in NhV deposition.
Due to the higher mass transfer coefficient for in-lake retention of NO3",
decreasing NO3" deposition decreases NO3" leaching but lowers ANC
production.
As SO42" and NO3- deposition decrease, a significant lake DOC
concentration increase has been observed. However, PnET-BGC does
not consider the decrease in soil organic matter partitioning as acid
deposition decreases and soil pH increases.
Zhou etal. (2015c)
MAGIC = ; NuCM = Nutrient Cycling Model; S042 = sulfate; Ca = calcium; Mg = magnesium; K = Potassium; S = sulfur; Be = base cation.
June 2018
4-102
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
4.6.2.
Critical Loads
4.6.2.1. Soil Acidification
The 2008 ISA included (McNultv et al.. 2007) a national assessment of soil critical loads
for acidification calculated by a SMBE technique (see Section 4.5.1.8 of the 2008 ISA).
The uncertainty associated with the SMBE technique was discussed in a second
publication (Li and McNultv. 2007). In general, the SMBE predicted that 26% of U.S.
forest soils have critical loads less than 1,000 eq/ha/yr. Low critical loads outside of New
England, New York, and the Appalachian Mountain region are not necessarily
problematic to forest health because acid deposition is much lower across most of the
U.S. compared to these areas. Unfortunately, mountainous terrain comprises much of the
area with very low forest soil critical loads. Mountain forests receive some of the highest
local rates of acidic deposition.
No new publications are identified in this review on the subject of national-scale
sensitivity; however, there are new reports on regional sensitivity. Figure 4-10 is a map of
soil CLs presented by McNultv et al. (2007) and updated with newer SMB modeling,
where available (McDonnell et al.. 2014b; Phelan et al.. 2014; Duarte et al.. 2013;
Sullivan et al.. 201 lb; Sullivan et al.. 201 la). Duarte et al. (2013) is a new evaluation of
critical loads for terrestrial acidification in New England, U.S. The steady-state mass
balance method is applied at over 4,000 plots. The acceptable ANC leaching rate was
calculated based on the critical chemical criteria of no change in base saturation. Over
80% of the critical loads were between 850 and 2,015 eq/ha/yr. Phelan et al. (2014)
evaluated PROFILE using national data sets as a method to estimate BCw rates for
forests in the U.S., focusing on Pennsylvania as the first test state. The model paired with
national data sets was successfully applied at 51 forested sites across Pennsylvania. The
soil critical loads were evaluated using a Bc:Al value of 10.0. The CL values ranged from
4 to 10,503 eq/ha/yr. Sullivan et al. (201 la) used the dynamic model MAGIC to model
terrestrial soil acidification in the Adirondacks, NY. Simulations were based on one
driver of acidic deposition (S) and included evaluation of CLs for soil solution molar
Bc:Al and Ca:Al ratio, two critical threshold levels (1 and 10), and two endpoint years of
model simulation (2050 and 2100). Statistically selected lakes (n = 44) were modeled to
represent a population of 1,320 sites. Nearly all (>93%) had a very low TL (<25 eq/ha/yr
for the Year 2100) when the protection threshold was set to BC:A1 = 10, and the majority
(>60%) had a high TL of (>100 eq/ha/yr for the Year 2100) to achieve Bc:AL = 1.
Sullivan et al. (201 lb) calculated surface water CLs for 66 sites in the Blue Ridge
mountains (North Carolina, Tennessee, and South Carolina). CL were reported for
June 2018
4-103
DRAFT: Do Not Cite or Quote

-------
1	surface waters; however, the MAGIC model was parameterized for all 66 sites including
2	terrestrial geochemistry. McDonnell et al. (2014b; studied acidification in the southern
3	Appalachians from northern Georgia to southern Pennsylvania and from eastern
4	Kentucky and Tennessee to central Virginia and western North Carolina. Although soil
5	critical loads were not reported in the publication, soil solution data are reported in
6	Figure 4-10.
Forest Ecosystems Critical Loads for Acidity
S + N eq/ha/yr
IH 170-1,000
¦I 1,001-2,000
; 2,001-4,000
	| 4,001-6,000
| 6,001-8,800
States
No Data
eq = equivalent; ha = hectare; yr = year.
(A.) McNultv et al. (2007) critical loads are mapped at 1-km2 grids (center map). For uncertainty, see Li and McNultv (2007).
(B.) Duarte et al. (2013) critical loads are mapped at 4-km2 grids; (C. and D.) Phelan et al. (2014) critical loads are mapped for each
sampling site (Pennsylvania). McDonnell et al. (2014b): Sullivan et al. (2011b): Sullivan et al. (2011a) critical loads are mapped as a
single point at the center point of the watershed (New York and North Carolina).
Source: http://nadp.sws.uiuc.edu/committees/clad/db/NCLDMapSummarv 2015.pdf.
Figure 4-10 Forest ecosystem critical loads for soil acidity related to base
cation soil indicators.
June 2018
4-104
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
In the case of semiarid ecosystems of southern California, which are not highly sensitive
to acidification, soils in high N deposition areas of chaparral and forested areas in the Los
Angeles basin have acidified significantly (Pardo et al.. 2011b).
4.6.2.2. Nitrate Leaching
Pardo etal. (2011b) documented the threshold N deposition value which caused
increased NOs leaching from forest ecosystems into surface water was 8 to
25 kg N/ha/yr. This information is summarized by ecoregion (Omernick Level 1) in
Figure 4-11. At 4 kg N/ha/yr, increasing NO3 was reported in the organic horizon in the
Colorado Front Range, which suggests incipient N saturation. In the northeastern U.S., N
budgets from 83 forested watersheds show that N retention averages 76% of incoming
atmospheric-N deposition and decreases from 90% retention for sites receiving
7 kg N/ha/yr to 60% retention for sites receiving 11 kg N/ha/yr (Aber etal.. 2003). The
highest critical loads were reported for Mediterranean California mixed-conifer forests.
Since the publication of Pardo etal. (2011b). new studies have been published on nitrate
leaching in the U.S. (see Appendix 4.3.2); one critical load is from Bowman et al. (2014).
who identify 10 kg N/ha/yr as the deposition level associated with nitrate leaching in
RMNP. New evidence from Europe by Pise et al. (2009) shows approximately 95% of
forests receiving less than 8 kg N/ha/yr have leaching of less than 1 kg N/ha/yr. In
Sweden, Khalili et al. (2010) showed a clear sudden increase in NO3 leaching in regions
where N deposition exceeded 7.5 kg/ha/yr.
June 2018
4-105
DRAFT: Do Not Cite or Quote

-------
Empirical CL of N (kg ha y )
4-17 Northwest Forested Mountains
8 Northern Forests; Eastern Temperate Forests
| 10-17 Mediterranean California
10-25 Great Plains
CL = critical load; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
The range of critical loads based on increased nitrate leaching for each ecoregion. The hatch marks indicate increasing levels of
uncertainty: no hatch marks for the most certain "reliable" category and single hatching for the "fairly reliable" category. White areas
lack data for critical loads determination for nitrate leaching.
Source: Pardo et al. (2011b').
Figure 4-11 Map of critical loads for nitrate leaching by ecoregion in the U.S.
4.7. Modification of Terrestrial Soil Response to Nitrogen (N)
Biogeochemical responses to N deposition can be modified by many environmental
factors including phosphorus, disturbance, stand age and climatic shifts in temperature
and precipitation. Here we provide a very brief overview of these topics. Appendix 13
provides an overview of modification of ecosystem response to N driven by climate,
whereas this section describes a brief summary of how climate modifies terrestrial soil
response to N.
Uncertainty
| Reliable
[VI Fairly Reliable
K>j Expert Judgment
June 2018
4-106
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
4.7.1. Disturbance and Stand Age Effects on Nitrogen Retention
The 2008 ISA reported that varying degrees ofN assimilation, leaching, and microbial
transformation often reflect differences in N status among treatment sites. These
variations have most often been attributed to disturbance history, dating back a century or
more (Goodalc and Aber. 2001). Sites that have undergone disturbances that cause loss of
soil N, such as logging, fire, and agriculture, tend to be most effective at retaining
atmospheric and experimental inputs of N. Fire causes substantial N losses from
ecosystems. Timber harvest contributes to nutrient removal from the ecosystem via
biomass export and acceleration of leaching losses (Mann ct al.. 1988; Bormann et al..
1968). In particular, logging contributes to loss of N and Ca2+ from the soil (Lattv et al..
2004; Tritton et al.. 1987). N retention capability often decreases with stand age,
suggesting that older forests are more susceptible than younger forests to becoming N
saturated (Hcdin ct al.. 1995).
One new study has been published since 2008 on how disturbance affects N retention
(Vourlitis and Pasquini. 2008). Fire did little to alter patterns of soil N enrichment from
atmospheric N deposition; however, periodic fires have important implications for the
structure and function of chaparral shrublands and their propensity to become N saturated
under current and future N deposition and fire regimes.
4.7.2. N and P Interactions
Mechanisms driving terrestrial N vs. P (phosphorus) limitations differ greatly since their
sources and biogeochemical dynamics differ; therefore, there is variability across
ecosystems (Vitousck et al.. 2010). Organic N can be decomposed by a variety of
enzymes to mineralize N from substrate; however, organic P relies on phosphatase
enzymes to mineralize it, independent of C respiration and N mineralization (Marklcin et
al.. 2016).
Literature on N deposition's effect on N and P dynamics focuses on N and P cycling and
the influence of N on phosphatase enzyme activity. Phosphatase is an enzyme released
into the soil by organisms. It cleaves or hydrolyzes phosphorous (phosphoric acid) from
substrate (soil, roots) to yield available phosphate for uptake (Marklein et al.. 2016). The
chemical structure of phosphatase includes N, therefore
Vitousek et al. (2010) explain that N fertilization increases P cycling because it enables
organisms to increase their extracellular phosphatase and, in turn, release more phosphate
from soil organic matter.
June 2018
4-107
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
There are some conflicting ideas of how N addition effects P in soils. N addition can
acidify soil, therefore phosphatase activity may be suppressed by this mechanism,
decreasing P release to soil (Wang et al.. 2016b). In contrast, N addition can decrease the
adsorption capacity of P ions in the soil, thus increasing P availability and uptake by
plants for productivity. It is also well known that N addition can increase increased litter
decomposition, which may also increase P release/cycling rates. An increased supply of P
can delay the onset of P limitation. It is important though to acknowledge that the type of
N applied, application rate, and ecosystem type may influence the phosphatase activity
(Song et al.. 2017).
4.7.3. Climate Modification of Acidification Effects on Soil
For soil acidification, potential future changes in the quantity and temporal distribution of
precipitation and temperature (and their interactions) is expected to alter the wet-dry
cycles that govern the timing and amount of acidic inputs in precipitation, microbial
transformation in the soil, and the flush of acid anions from soils to surface waters.
There are two recent papers on the relationship between precipitation and sulfur in U.S.
watersheds. Rice et al. (2014) indicate that hydrology, and specifically runoff, is an
important controller of sulfate recovery in watersheds because drainage flushes the
accumulated sulfur from the soil. If precipitation and runoff patterns change under a
future climate, this important process will be affected. Similarly, Mitchell et al. (2011)
observed that following decades of changes in stream sulfate concentrations and fluxes
that have been driven by atmospheric deposition, variation is stream sulfate in now being
controlled by variations in precipitation inputs. Increased variation in precipitation will
increase wetting and drying cycles that promote mineralization of sulfate from soil and
subsequent methylation.
In general, if acid anions build up in soil during periods of drought, the eventual flushing
likely causes a more potent acidification event (Moslev. 2015; Whitehead et al.. 2009). If
the acidification event occurs during a time when sensitive biota or lifestages of biota are
present, acidification may cause more adversity to these populations (Kowalik et al..
2007). Increases in storm frequency associated with global climate change (Collins et al..
2014) could increase the frequency and severity of acidification driven by high levels of
sea salt deposition in coastal regions (Wright and Schindler. 1995). Although the
mechanisms of interaction are unclear, increases in DOC concentrations in aquatic
ecosystems across Europe and the U.S. have been linked to acidification, N cycling, and
climate change, with important implications for water quality and ecosystem function
(Evans et al.. 2008) (see Appendix 4.3.9).
June 2018
4-108
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Warmer temperatures increase decomposition and nitrification. Nitrification will also
increase with increased N supply caused by increased weathering or decomposition
(Booth et al.. 2005). The process of nitrification generates protons that increase the rate
of nitrate and base cation leaching to drainage waters (Murdoch et al.. 1998). The
combined increase of NO3 leaching and loss of base cations has the potential to magnify
acidification in forest soils (Fernandez et al.. 2003). Soil weathering is typically the key
buffer to acidic deposition (Li and McNultv. 2007). and while weathering is increased by
both soil temperature and soil moisture (Gwiazda and Broecker. 1994). it is unclear
whether any future change in the magnitude of temperature and precipitation will be
enough to alter base cation supply or influence the acid-base balance of sensitive
ecosystems. Furthermore, it is unclear whether increased supply of N in soils from either
deposition, increased decomposition, or increased nitrogen fixation will negate the
ameliorative effect of enhanced weathering. Some studies show that climate change will
mitigate acidification through increased weathering (Belvazid et al.. 201 la), while others
show that climate change will aggravate acidification although increased nitrification
outpacing enhanced weathering (Wu and Driscoll. 2010). In general, increased
temperature and precipitation will likely enhance inputs of buffering agents from
weathering and deposition, but also increase inputs of acidifying agents from deposition
and enhanced N cycling. The relative sensitivity of these opposing processes to a given
change in climate remains unresolved.
Ecological effects of atmospheric nitrogen (N) and sulfur (S) deposition on two
hardwood forest sites in the eastern United States were simulated in the context of a
changing climate using the dynamic coupled biogeochemical/ecological model chain
ForSAFE-Veg. The main driver of ecological effects was soil solution N concentration
(Mcdonnell et al.. 2018a).
4.7.4. Climate Modification of Nitrogen-Driven Eutrophication in Soil
The following is brief summary of how temperature, snow, and precipitation affect soil
response to N. In addition, information is summarized on N effects on the flux of three
GHGs (CO2, CH4 and N2O) from terrestrial ecosystems.
Snow interacts with soil biogeochemistry in a number of ways, mostly by temperature
shifts associated with snowpack and soil moisture flux during snowmelt. The highest
flow and lowest ANC periods of the year tend to coincide with times of high snowmelt.
This is when acidification is likely to occur. Snow pack insulates soils from frigid winter
air temperatures. With a warming climate, soils are expected to become colder and
experience more winter soil freeze-thaw cycles as snow cover continues to decline.
June 2018
4-109
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
A number of experiments have evaluated snow effects on soil N; these include studies on
the timing, frequency, and severity (snow depth). Groffman et al. (2001) conducted a
snow manipulation experiment at HBEF and found mild late-season freezing increases
soil N03 levels by physical disruption (increased fine root mortality) causing reduced
plant N uptake and reduced competition for inorganic N, allowing soil NOs levels to
increase even with no increase in net mineralization or nitrification. Later studies at the
same site have shown that soil N mineralization rates were more strongly related to soil
volumetric water content than to root biomass, snow or soil frost, or winter soil
temperature (Sorensen et al.. 2016). and the more freeze/thaw cycles anticipated with
climate warming supported lower rates of N mineralization at HBEF (Duran et al.. 2016).
The effects of snow depth on soil N have been synthesized in a meta-analysis by Li et al.
(2016c). who evaluated the central tendencies from 41 publications on 12 variables
related to soil N. Results are that snow depth significantly increased foliar N (+4.5%) and
microbial N (MBN, +35.9%), but significantly decreased soil N2O efflux (-34.1%) and
nitrification (-24.8%). However, altered snow depth did not significantly affect soil
dissolved organic N (DON), total inorganic N, nitrate, and N leaching. Because
increasing snowpack depth promoted MBN, the unchanged, net N mineralization and soil
ammonium content were probably due to limitation of the soil N availability and other
soil abiotic factors rather than soil microbes.
Greaver et al. (2016) created a synthesis of meta-analyses to study the effects of the
individual effects of either N, temperature, or precipitation on different pools and
processes related to ecosystem C pools and fluxes (Figure 13 2). Nineteen different pools
and processes were included in this analysis.
Initial findings are consistent with the biokinetic effects of warming; long-term data and
meta-analyses show that soil respiration, including decomposition and microbial
respiration, is stimulated by increasing temperature (Lu et al.. 2013; Bond-Lambertv and
Thomson. 2010; Rustad et al.. 2001). Most empirical studies show rising temperature
stimulates N release by mineralization (Churkina et al.. 2010). which may be driven more
by temperature effects on moisture (Emmett et al.. 2004). In some dynamic land models,
the additional N from mineralization will stimulate C uptake by plants even more than
current N deposition (Burd et al.. 2016). At the same time, increased N from
mineralization may cause N induced inhibition of decomposition, a feedback mechanism
that might decrease the amount of N released and which is currently considered by a few
models (Gerberetal.. 2010). The mechanisms causing N driven reduction in
decomposition are not well understood but are thought to result from changes in
microbial community composition and the production of decomposition enzymes, as well
as possible changes in the character and degradability of soil organic matter (Conant et
al.. 2011; Janssens et al.. 2010). Climate change could also affect decomposition rates by
4-110
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
altering both available soil moisture and microscale connectivity among microorganisms,
water, and nutrients within the soil matrix, which, in turn, may alter microbial processes
(Xiang et al.. 2008). Although there is no consensus about how dissolved organic carbon
(DOC) in surface water is regulated overall, increasing N and temperature increase DOC
concentrations (Laudon et al.. 2012). While few meta-analyses have examined
precipitation effects on the soil C cycle, precipitation tends to increase the root C pool
(Figure 13-2).
Ni et al. (2017) created a meta-analysis to evaluate the interactions between warming and
N deposition and found the interaction greatly increased the soil C input (+49%)
compared with the single factor of either warming (+5%) or N deposition (+20%). Soil C
loss was not significantly affected by the interaction of N and warming likely because
increases in decomposition due to warming are offset by the decreases by N addition.
The N addition alters fluxes of GHGs by regulating plant and microbial activities that are
directly associated with GHG production and consumption processes (Figure 4-12). It is
extremely important to consider total ecosystem flux because consumption may be offset
by production, even within the same ecosystem. For example, N addition stimulates plant
growth in most ecosystems (LeBauer and Treseder. 2008). and may in turn increase C
sequestration in plant biomass. On the other hand, maintenance respiration is positively
correlated with tissue N content (Reich et al.. 2008). and litter with higher N content also
decomposes faster. Therefore, increased leaf N content under elevated N may result in
higher C loss by increasing both autotrophic and heterotrophic respiration. Like CO2,
ecosystems may consume and produce CH4 and N2O, and the balance determines their
net release.
Liu and Greaver (2009) conducted a meta-analysis of 313 observations across 109 studies
to evaluate the effect of N addition on the flux of three GHGs: CO2, CH4 and N2O. The
results indicated that N addition increased ecosystem carbon content of forests by 6%,
marginally increased soil organic carbon of agricultural systems by 2%, but had no
significant effect on net ecosystem CO2 exchange for nonforest natural ecosystems.
Across all ecosystems, N addition increased CH4 emission by 97%, reduced CH4 uptake
by 38%, and increased N2O emission by 216%. Most often, N addition is considered to
increase forest C sequestration without consideration of N stimulation of GHG
production in other ecosystems. However, this study indicated that although N addition
increased the global terrestrial C sink, the CO2 reduction could be largely offset
(53-76%) by N stimulation of global CH4 and N2O emission from multiple ecosystems.
4-111
DRAFT: Do Not Cite or Quote

-------
CO
n2o
*
N Input
Denitrifying bacteria (*¦)

n2^np^no\
s. Anaerobic
NH4 _>NH,ONO —NO,
" ^
NjO
NOj
Nitrifying bacteria (+)
Aerobe
SOC

DOC

DIN/DON
C cycle
N cycle
+ Positive feedback
- Negative feedback
ANPP = aboveground net primary productivity; BNPP = belowgrourid net primary productivity; SOC = soil organic carbon;
DOC = dissolved organic carbon; DIN = dissolved inorganic nitrogen; DON = dissolved organic nitrogen.
Source: Liu and Greaver (2009).
Figure 4-12 The potential mechanisms that regulate the responses of carbon
dioxide (CO2), methane (CH4), and nitrous oxide (N2O) production
and consumption to elevated nitrogen (N).
4.8. Summary
4.8.1.
Sources
The effects of N and S deposition on soil biogeochemistry cause cascading effects on
biological species. The biological effects are discussed in Appendix 5 and Appendix 6.
Since 2008, there have been and number of estimates of sulfate, oxidized nitrogen, and
reduced nitrogen from atmospheric deposition. The most recent estimates are
summarized in Appendix 2. and maps showing the geographic distribution of deposition
are presented for total acidifying deposition (Figure 2-12). total N deposition
(Figure 2-13). and total S deposition (Figure 2-31). Maps depicting how the relative
June 2018
4-112
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
contribution of oxidized and reduced N species varies across the U.S. are presented in in
Figure 2-14.
In the 2008 ISA, atmospheric deposition was identified as the main source of
anthropogenic N, relative to other N sources, to nonmanaged terrestrial ecosystems. This
has been confirmed by new studies on N sources to U.S. lands and waterways, which find
human-mediated N inputs are spatially heterogeneous across the country, ranging from
<1.0 to 34.6-fold the rate of background N input across the conterminous U.S. Synthetic
N fertilizer and atmospheric N deposition are the largest and second-largest overall
human-mediated N sources (the single largest sources in 41 and 33% of HUC-8s,
respectively). There is no new information published on non- deposition sources of S in
terrestrial ecosystems. S inputs from emissions to the atmosphere are discussed in
Appendix 2.
4.8.2. Soil Processes and Indicators
Soil N enrichment and soil acidification are occuring in sensitive ecosystems across the
U.S. at recent levels of deposition. Soil acidification is a natural process that can be
accelerated by N and/or S deposition. A number of soil geochemical processes are
associated with acidification (Appendix 43 and Table 4-19). Base cations counterbalance
acid anions. Base cations are added to a watershed by weathering and atmospheric
deposition and are removed by leaching and (perhaps temporarily) by uptake into
growing vegetation. Acidifying deposition can deplete exchangeable base cation pools in
soils where acid deposition is high relative to base cation input. The limited mobility of
anions associated with naturally derived acidity (organic acids and carbonic acid)
controls the rate of base cation leaching from soil where rates of atmospheric deposition
of S and N are low. In addition to increasing ion concentration in soil, inputs of S and N
in acidifying deposition provide anions that are highly mobile, further accelerating the
rates of base cation leaching. In addition, the deposition of reduced forms of N
(e.g., NHx) can stimulate nitrification, which is the microbial oxidation of NH4+ to NOs .
This oxidation process acidifies soils because 2 moles of hydrogen ion (H+) are released
per mole of NH4+ converted to NO3 . Therefore, both chemically reduced and oxidized
forms of N and S deposition contribute to terrestrial acidification. Several useful
indicators of soil acidification (Table 4-19) include indicator thresholds related to
biological responses, the biological basis of which are discussed in Appendix 5.
It is clear from Table 4-19 that some of the same processes and indicators associated with
acidification are also associated with N enrichment. The N enrichment of terrestrial soils
occurs in response to the input of exogenous N. The new studies on N deposition
June 2018
4-113
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
contribution to total N loading in terrestrial U.S. ecosystems are discussed in
Appendix 42.
The 2008 ISA documented that most N, often more than 85% of the total ecosystem N, is
stored in the soil in temperate forest ecosystems. There is new evidence that soil litter is
the largest N pool in grasslands, shrublands, and wetlands (Appendix 4.3. IV The ability
of atmospheric deposition to increase soil N is indicated by a positive correlation between
atmospheric deposition levels and total N concentration in the upper soil horizon. Soil N
accumulation is linked to increased N leaching and decreased N retention
(Appendix 4.3.2V Thresholds of N deposition that are associated with the onset of
elevated NOs leaching are discussed in Appendix 4.6.2.2.
Recent work suggests that N leaching from soil can precede the complete saturation of
the biotic and abiotic processes that retain N (Appendix 4.3.2). Lovett and Goodale
(2011) proposed a model of N saturation in which NO3 leaching can occur even if the
ecosystem N retention capacity has not yet been saturated, as is observed at many sites
and supported by several new studies. Although N leaching may occur prior to saturation,
N budgets from 83 forested watersheds in the northeastern U.S. show that N retention
averages 76% of incoming N deposition. Nitrogen retention decreases from 90%
retention for sites receiving 7 kg N/ha/yr to 60% retention for sites receiving
11 kg N/ha/yr, suggesting a range of response in which leaching is exacerbated in this
region (Aberetal.. 2003).
The 2008 ISA documented that N enrichment alters rates of microbial transformation of
chemicals in the soil, such as increased rates of nitrification and denitrification, and
altered rates of decomposition. The addition of N to terrestrial ecosystems can increase
nitrification, an important process for soil acidification and the production of NO, .
Nitrification is often stimulated in soils with a C:N ratio below approximately 20 to 25,
and N deposition often causes the ratio to fall below this threshold. The NO3 created by
nitrification may be leached or denitrified. In terrestrial ecosystems, denitrification of
N03 mainly occurs in saturated zones. Several syntheses have been published since 2008
evaluating N addition effects on denitrification and nitrification in terrestrial ecosystems
(Appendix 4.3.6). A new meta-analysis shows N addition significantly increases
denitrification from many types of ecosystems (coniferous forest, deciduous forest,
tropical forest, wetland, grassland), except heathlands.
4-114
DRAFT: Do Not Cite or Quote

-------
Table 4-19 Summary of key soil geochemical processes and indicators associated with eutrophication and
acidification.
Endpoint
N-Driven Nutrient
Enrichment
Acidification
The Effect of Deposition
PROCESS
N saturation
X
X
New empirical evidence suggests revising the N saturation concept, because
observed NO3" leaching can occur even if the ecosystem N retention capacity
has not yet been saturated.
Soil N accumulation
X
X
New meta-analysis across ecosystem types confirms inorganic soil NO3"
concentration increases with N addition. New gradient study confirms that N
concentration increases with N deposition. New addition study confirms
increased soil accumulation.
NO3" leaching
X
X
New meta-analysis confirms leaching increases with N addition.
New regional-scale gradient analyses: <8 kg N/ha/yr onset of leaching
<1 kg N/ha/yr in European forests. Second study suggests <7.5 kg N/ha/yr
increases leaching in Swedish forests.
New addition study: 9 to 14 kg N/ha/yr increase in inorganic N concentrations.
New USFS critical loads for the onset of leaching: 8 to 10 kg N/ha/yr in eastern
and western U.S., 17 kg N/ha/yr in the Sierra Nevada and San Bernardino
Mountains.
S accumulation and
adsorption

X
Some soils (notably in many watersheds in the SE U.S.) have the capacity to
adsorb substantial quantities of S, with essentially no acidification of drainage
water. Nevertheless, S adsorption capacity is finite and under continual high S
deposition loading, the adsorptive capacity of soil will eventually be depleted.
New studies of 27 watersheds in the SE indicates most will begin releasing
SC>42"in the next two decades, and northeastern watersheds show a net loss of
S from soils now in response to decreased levels of atmospheric S deposition.
SO42" leaching

X
Atmospheric S deposition generally increases leaching of SO42" to surface
waters. The amount of deposition that causes the onset of leaching varies
across the landscape.
June 2018
4-115
DRAFT: Do Not Cite or Quote

-------
Table 4-19 (Continued): Summary of key soil geochemical processes and indicators associated with
eutrophication and acidification.
N-Driven Nutrient
Endpoint	Enrichment	Acidification
Base cation	X
release/depletion
The Effect of Deposition
Base cation (Ca, Mg, K, Na) release occurs in response to the input of acid
anions (SO42" and NO3") from deposition. This proportional change in base
cations relative to acid anions from N and S deposition is called the F-factor.
New studies confirm base cation depletion continues to occur in the Rocky
Mountains (threshold 28 kg N/ha/yr) and in U.K. grasslands, while in a
northeastern forest, 17 yr of N addition did not cause further depletion. A
meta-analysis indicates cation depletion occurs early after increased deposition
of acid anions, which tapers off with time.
Al mobilization	X	<15 to 20% soil base saturation is the threshold for inorganic Al mobilization
from soil. This is an extremely important effect of acidifying deposition because
inorganic monomeric Al, including Al3+ and various hydroxide species, is toxic to
biota (Appendix 5). Inorganic Al is minimally soluble at pH about 6.0, but
solubility increases steeply at pH values below about 5.5.
There have been no new studies on Al mobilization.
Nitrification	X	X	Nitrification releases 2 mol hydrogen ion (H+) per mol NHV converted to NO3",
acidifying soils. As soil inorganic N accumulates, net nitrification rates often
increase, and NO3" can leach from the ecosystem.
New N gradient and meta-analysis studies confirm N addition increases
nitrification.
Denitrification	X	Denitrification is the microbial reduction of NO3" to nitrite (NO2"), nitric oxide
(NO), the greenhouse gas nitrous oxide (N2O), and N2. It occurs under
anaerobic conditions. In Europe, soil switched from a source to a sink after two
decades of N deposition exclusion.
New meta-analysis confirms N addition increases denitrification rates.
DOC leaching	X	X	Acidity of some surface waters is partly regulated by DOC concentrations. Soil
acidification can suppress the natural production of DOC. In recent years, the
DOC concentration of some lakes and streams has risen, with adjacent
terrestrial ecosystems as the DOC source. However, the mechanism for this
increase is unclear. It may be due to soil recovery from acidification or N
deposition effects on decomposition.
New studies include a meta-analysis and field addition studies.
June 2018
4-116
DRAFT: Do Not Cite or Quote

-------
Table 4-19 (Continued): Summary of key soil geochemical processes and indicators associated with
eutrophication and acidification.
Endpoint
N-Driven Nutrient
Enrichment
Acidification
The Effect of Deposition
Decomposition/
mineralization
X
X
Decomposition rates correlate with ratios of C:N, lignin:N, or lignin:cellulose in
litter. The addition of N can stimulate the decomposition of labile compounds
that degrade during the initial stages of decomposition, but added N can
suppress the decomposition of more recalcitrant material. Evidence for this is
widespread in forests but has not yet been well documented in grasslands and
other ecosystems.
There are new addition studies and meta-analysis to better understand the
mechanisms and response trends.
INDICATOR
Soil [N]
X
X
Increasing N deposition increases concentration of N in soil and reflects soil N
accumulation that may be assimilated by organisms or mobilized via leaching.
Soil C:N ratio
X
X
Increasing N deposition decreases the C:N in plant tissue, causing cascading
effects during decomposition on microbial transformation, especially nitrification,
leading to increased NO3" leaching.
THRESHOLD: <20 to 25 associated with elevated rates of nitrification and
elevated risk of nitrate leaching in the U.S. and <25 to 30 for increased NO3"
leaching in Europe.
Soil base saturation

X
Increasing N + S deposition decreases the soil pool of exchangeable base
cations.
THRESHOLD: <15 to 20% is the value at which exchange ion chemistry is
dominated bv inoraanic Al and mav cause iniurv to veaetation (see Appendix 5).
Soil Bc:AI ratio	X	Increasing N + S deposition decreases the soil pool of exchangeable base
cations, often decreasing the Ca:AI ratio.
THRESHOLD: Ca:AI <1.0 to 10 causes physiological stress and decreased
growth and survival of sensitive plant species (see Appendix 5).
Fungi:bacteria ratio	X	New indicator since 2008 ISA: Increasing N deposition decreases the fungi-to-
bacteria ratio and causes a transition from N to C limitation among soil food
webs.
Al = aluminum; Be = base cation; C = carbon; Ca = calcium; DOC = dissolved organic carbon; H+ = hydrogen ion; ha = hectare; K = potassium; kg =kilogram; Mg = magnesium;
N = nitrogen; N2 = molecular nitrogen; N20 = nitrous oxide; NE = northeast; NH4+ = ammonium; NO = nitric oxide; N02" = nitrite; N03" = nitrate; S = sulfur; SE = southeast; yr = year.
June 2018
4-117
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
The effects of N on decomposition, which is the breakdown of organic matter, is an
active area of research (Appendix 4.3.7V Within the soil microbial community, bacteria
and fungi are the primary decomposers of organic matter. Both microbial community
composition (see Appendix 6) and microbial enzyme activity can respond dynamically to
shifts in inorganic nutrient and substrate availability, reflecting the nutrient and energy
limitation of the microbial community. Litter decay rates are also well established to
correlate with ratios of C:N, lignin:N, or lignin:cellulose in litter. These chemical traits
have been shown to account for over 73% of the variation in litter decomposition rates
worldwide. Based on these observations, it could be assumed that added N would
stimulate decomposition and the loss of C from soil pools. There is now widespread
evidence, however, that the stimulatory effects of N on decomposition are limited to the
early stages of mass loss, when more labile compounds are degraded. During the later
stages of decomposition when the rates of mass loss are slow and controlled by the
degradation of recalcitrant compounds, the addition of N slows decomposition. This
slowing of decomposition appears to be a consequence of decreases in the production of
some extracellular enzymes by fungi. It is clear that N additions generally decrease
respiration from soil heterotrophs, but this may not decrease total soil respiration because
heterotrophic respiration accounts for only a portion of soil CO2 efflux.
Several new meta-analyses have been published since 2008 on N addition effects on
belowground carbon cycling (Appendix 4.3.10'). About half of C fixed annually by
terrestrial vegetation is allocated to belowground C pools; therefore, it is important to
understand how N affects belowground C to better understand changes in plant
physiology, plant growth, and ecosystem C cycling (Appendix 6). Many studies have
shown that the belowground C cycle does not always match the aboveground C cycle.
Therefore, it is inappropriate to extrapolate from aboveground responses to belowground
processes.
4.8.3. Monitoring
Several new studies use long-term monitoring data sets. (Yanai et al.. 2013) evaluated
45 years of biogeochemical monitoring data at the Hubbard Brook Experimental Forest
LTER, New Hampshire, U.S. Since 1992, the ecosystem shifted to a net N sink of
~8 kg N/ha/yr. Mitchell and Likens (2011) examines sulfur accumulation observed in
over four decades of continuous long-term record for four watersheds (reported in
Appendix 4.3.3). At the Niwot Ridge LTER, CO, U.S., Lieb et al. (2011) found that a
decade of simulated N deposition to alpine ecosystems caused ongoing changes in
diversity and soil biogeochemistry, including lower soil acid buffering capacity,
decreased concentrations of Mg2+, and increased concentrations of the potentially toxic
June 2018
4-118
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
cations Mn2+ and Al3+. Their results suggested an N deposition threshold of
28 kg N/ha/yr.
There are new studies on long-term monitoring in Europe, including investigations of S
dynamics in England (Lchmann et al.. 2008) and N and S dynamics in Switzerland
(Pannatier ct al.. 2011) and N and C interactions between boreal soils and lakes in
Sweden, (Khalili et al.. 2010). Notably, Khalili et al. (2010) found a significant relation
between C:N ratios of the organic soil layer and C in lake waters, and a clear sudden
increase in NOs leaching regions where N deposition exceeded 7.5 kg N/ha/yr.
4.8.4. Models
The most commonly used ecosystem models in the U.S. were described in the 2008 ISA.
Here we focus on the several primary models that are currently being used in the U.S. to
assess the effects of S and N deposition on terrestrial ecosystems. One important input to
these models are estimates of base cation weathering (BCw). There are new updates on
two methods to estimate this parameter: Soil Texture Approximations (STA) and
PROFILE. Steady-state models include steady-state mass balance equations (SMBE),
while dynamic models include the VSD and VSD+, MAGIC, NuCM, PnET/BGC, and
DayCent-Chem.
Base cation weathering rate is one of the most influential yet difficult to estimate
parameters in the calculation of critical acid loads of nitrogen (N) and sulfur (S)
deposition for terrestrial systems. Commonly used models include STA, a simple
empirical steady-state model with low data requirements, and PROFILE, a mechanistic,
steady-state kinetics model that requires 26 input parameters. There is a new study on
estimating BCw. Koseva et al. (2010) confirm that the STA equations were derived for
young soils that developed following the Late Wisconsin glaciation and may not be
suitable for older, more weathered soils that were not affected by the most recent
glaciation and which cover the majority of the U.S. (U.S. EPA. 2009c). The larger data
set required to run PROFILE prohibits applying the model to many parts of the U.S.
However, a recent soil geochemical data set compiled by the USGS has made it possible
to apply PROFILE in the U.S. The initial application in 51 forested sites across
Pennsylvania indicates this technique may be appropriate for applications nationwide
(Phelan et al.. 2014). The uncertainty associated with the calculation of BCw has recently
been evaluated by Futteretal. (2012). who estimated that the contribution of model
parameter uncertainty to overall variability related to PROFILE and MAGIC (a dynamic
model that typically uses PROFILE for initial estimates of BCw) was relatively small.
June 2018
4-119
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Estimates of BCw are input parameters in soil acidification models. The simplest
acidification models are the steady-state mass balance models. The 2008 ISA
documented a model based on simple mass-balance equations (SMBE) to assess critical
loads for acidification in U.S. forest soils (McNultv et al.. 2007). A second publication
discussed the uncertainties associated with this model and national-scale assessment (Li
and McNultv. 2007). The results indicated that uncertainty in the critical load came
primarily from components of BCw (49%) and acid neutralizing capacity (46%), whereas
the most critical parameters were BCw base rate (62%), soil depth (20%), and soil
temperature (11%). The authors concluded that improvements in estimates of these
factors are crucial to reducing uncertainty and successfully scaling up SMBE for national
assessments. Posch et al. (2011) discusses a regional application of SMB models.
Dynamic models are typically applied to much smaller spatial areas than steady-state
models because they require more data and parameterization. The VSD model is the
simplest extension of the steady-state SMB model into a dynamic model that is designed
for sites with few data available and applications on a large regional or continental scale.
Posch and Reinds (2009) have developed a version of the V SD for steady-state critical
load applications at the regional scale. The ForSAFE model Wallman et al. (2005)
simulates the cycles of carbon, nitrogen, base cations (Be), and water in a forest
ecosystem, while simultaneously simulating soil chemistry, tree growth, and soil organic
matter accumulation or depletion. ForSAFE-VEG is a composite model (Sverdrup et al..
2007) that can link changes in atmospheric deposition, climatic conditions, and land use
to responses in the biogeochemistry and plant community composition at the site level,
both historically and into the future. Belvazid et al. (201 la) stated that the
biogeochemical model platform must be improved to simulate N processes more
accurately before being applied to calculate critical loads. The model overestimated the
actual N concentrations in the soil solution. Several new applications of ForSAFE-VEG
were published [(McDonnell et al.. 2013; Sverdrup et al.. 2012; Belvazid et al.. 201 la),
and the results are discussed in Appendix 5 and Appendix 61.
MAGIC (Cosby etal.. 1985a; Cosby et al.. 1985c; Cosby etal.. 1985b) is one of the most
well-known dynamic models of aquatic acidification. An update to MAGIC by Oulehle et
al. (2012) gives a more realistic simulation of observed changes in N leaching. The new
formulation also provides a reasonable simulation of the long-term changes in C and N
pools (and C:N ratio) with SOM. PnET-BGC is an integrated forest-soil-water model that
has been used to assess the effects of air pollution and land disturbances on forest and
aquatic ecosystems; it uses a new conceptual advancement that incorporates CO2
dynamics into the model. DayCent-Chem links two widely accepted and tested models,
one of daily biogeochemistry for forest, grassland, cropland, and savanna systems,
DayCent (Parton et al.. 1998). and the other of soil and water geochemical equilibrium,
4-120
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
PHREEQC (Parkhurst and Appelo. 1999). There is a new conceptual advancement for
Day Cent which now includes O3 (Bvtnerow icz et al.. 2013). The Nutrient Cycling Model
(NuCM) was developed to synthesize current understanding of nutrient cycling in forests
and to predict how forests respond to changing S and N atmospheric deposition rates. No
revisions to these models have been identified.
4.8.5. National-Scale Sensitivity
The 2008 ISA documented that by the end of the 1980s, after nearly a decade of intensive
research performed under the original NAPAP, the regions of the U.S. with many
acid-sensitive waters and ecosystems were well recognized. These acid-sensitive
ecosystems are generally located in upland mountainous terrain in the eastern U.S. and
are underlain by weathering-resistant bedrock like granite or quartzite sandstone.
However, similar work characterizing areas sensitive to the eutrophication effects of
nitrogen is not available. Typically, all terrestrial ecosystems are sensitive to N
deposition.
There is strong evidence to show that biogeochemical sensitivity to N-driven
eutrophication and acidification from atmospheric deposition is the result of historical
loading, geologic/soil conditions (e.g., mineral weathering and S adsorption), and
nonanthropogenic sources of N and S loading to the system. No single deposition level is
applicable to all ecosystems in the U.S. that will describe the onset of eutrophication or
acidification. Ecosystem sensitivity is heterogeneous.
New publications have commented on recovery of terrestrial ecosystems from
acidification at the national scale (NAPAP. 2011). specifically in the Northeast (Phelan et
al.. 2016; Fuss et al.. 2015; Lawrence et al.. 2015a). and the lack of recovery in the
southern Appalachian Mountains (Fakhraci et al.. 2016; Zhou et al.. 2015b; McDonnell et
al.. 2013; Elliott et al.. 2008). Work on national-scale soil acidification published in 2007
remains the most recent national assessment of this effect (McNultv et al.. 2007);
however, Figure 4-10 is a map of soil CLs that updates (McNultv et al.. 2007) with newly
available modeling (McDonnell et al.. 2014b; Phelan et al.. 2014; Duarte et al.. 2013;
Sullivan et al.. 201 lb; Sullivan et al.. 201 la).
New critical loads for nitrate leaching in forest ecosystems by Pardo et al. (2011b) ranges
from 8 to 25 kg N/ha/yr across ecoregions (Omernick Level 1) in the U.S. (Figure 4-11).
At 4 kg N/ha/yr, increasing NO3 was reported in the organic horizon in the Colorado
Front Range, which suggests incipient N saturation. In the northeastern U.S., N budgets
from 83 forested watersheds show that N retention averages 76% of incoming
atmospheric-N deposition and decreases from 90% retention for sites receiving
June 2018
4-121
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
7 kg N/ha/yr to 60% retention for sites receiving 11 kg N/ha/yr (Aberetal.. 2003). The
highest critical loads were reported for Mediterranean California mixed-conifer forests.
Since the publication of Pardo etal. (2011b). new studies have been published on nitrate
leaching (see Appendix 4.3.2). Evidence from Europe by Pise et al. (2009) shows
approximately 95% of forests receiving less than 8 kg N/ha/yr have leaching of less than
1 kg N/ha/yr. In Sweden, Khalili et al. (2010) showed a clear sudden increase in NOs
leaching in regions where N deposition exceeded 7.5 kg/ha/yr.
4.8.6. Climate Modification of Soil Response to N Addition
Soil biogeochemistry response to N deposition can be modified by climatic shifts in
temperature and precipitation. For soil acidification, potential future changes in the
quantity and temporal distribution of precipitation and temperature (and their
interactions) is expected to alter the wet-dry cycles that govern the timing and amount of
acidic inputs in precipitation, microbial transformation in the soil, and the flush of acid
anions from soils to surface waters. In general, increased temperature and precipitation
will likely enhance inputs of buffering agents from weathering and deposition, but also
increase inputs of acidifying agents from deposition and enhanced N cycling. The relative
sensitivity of these opposing processes to a given change in climate remains unresolved.
Snow interacts with soil biogeochemistry in a number of ways, mostly by temperature
shifts associated with snowpack and soil moisture flux during snowmelt (Duran etal..
2016; Sorensen et al.. 2016; Groffman et al.. 2001). The highest flow and lowest ANC
periods of the year tend to coincide with times of high snowmelt. This is when
acidification is likely to occur. Snow pack insulates soils from frigid winter air
temperatures. With a warming climate soils are expected to become colder and
experience more winter soil freeze-thaw cycles as snow cover continues to decline. There
have been a number of experiments evaluating snow effects soil N, these include studies
on the timing, frequency, and severity (snow depth) (Li et al.. 2016c).
Greaver et al. (2016) created a synthesis of meta-analyses to study the effects of either N,
temperature, or precipitation on different pools and processes related to ecosystem C
responses to these variables and to gain some insight as to the direction and magnitude of
the change observed (Figure 13-2). Nineteen different pools and processes were included
in this analysis. Liu and Greaver (2009) conducted a meta-analysis of 313 observations
across 109 studies of terrestrial and wetland ecosystems to evaluate the effect of N
addition on the flux of three GHGs (CO2, CH4 and N2O). This study indicated that
although N addition increased the global terrestrial C sink, the CO2 reduction could be
largely offset (53-76%) by N stimulation of global CH4 and N2O emission from multiple
June 2018
4-122
DRAFT: Do Not Cite or Quote

-------
1	ecosystems. The authors noted that N addition alters fluxes of GHGs by regulating plant
2	and microbial activities that are directly associated with GHG production and
3	consumption processes. It is therefore extremely important to consider total ecosystem
4	flux of GHGs because consumption may be offset by production, even within the same
5	ecosystem.
June 2018
4-123
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
APPENDIX 5. BIOLOGICAL EFFECTS OF
TERRESTRIAL ACIDIFICATION
This appendix characterizes the biological effects of acidifying deposition of nitrogen (N)
and sulfur (S) to terrestrial ecosystems. Appendix 5_2 discusses effects on trees and
forests (Appendix 5.2.1). understory plants and grasslands (Appendix 5.2.2). lichens
(Appendix 5.2.3). soil biota (Appendix 5.2.4). and fauna (Appendix 5.2.5). The
characteristics, distribution and extent of sensitive ecosystems is presented in
Appendix 5.3. Next, Appendix 54 presents the application of terrestrial acidification
models. Finally, levels of deposition at which effects are manifested is discussed in
Appendix 5.5. including a discussion of critical loads (Appendix 5.5.3). Climate
modification of acidifying deposition effects is discussed in Appendix 5.6. A summary
section with causal determinations based on a synthesis of new information and previous
evidence of biological effects of terrestrial acidification is presented in Appendix 5/7.
5.1. Introduction
Changes in biogeochemical processes caused by acidifying deposition of N and S to
terrestrial systems (Appendix 4) are linked to changes in terrestrial biota and have
significant ramifications for biological functioning of these ecosystems. In the 2008 ISA
for Oxides of Nitrogen and Oxides of Sulfur—Ecological Criteria (hereafter referred to as
the 2008 ISA), the evidence was sufficient to infer a causal relationship between
acidifying N and S deposition and changes in terrestrial biota. The strongest evidence for
a causal relationship comes from studies of terrestrial systems exposed to elevated levels
of acidifying deposition that show decreased plant health, reduced plant vigor, and/or loss
of terrestrial biodiversity. In multiple studies, consistent and coherent evidence has
shown that acidifying deposition can affect terrestrial ecosystems by causing direct
effects on plant foliage and indirect effects associated with changes in soil chemistry
(Section 3.2.2.3 of 2008 ISA). Biological indicators with reported responses to acidifying
deposition and conditions created by acidifying deposition include forest trees, shrubs,
lichens, grasslands, and Arctic and alpine tundra.
Acidifying deposition can affect terrestrial ecosystems via direct effects on plant foliage
and indirect effects associated with changes in soil chemistry (Figure 5-1). Biological
effects of acidification on terrestrial ecosystems are generally attributable to aluminum
(Al) toxicity and decreased ability of plant roots to take up base cations (especially
calcium [Ca]) and water from the soil (Cronan and Grigal. 1995). Acidifying deposition
to acid-sensitive soils can cause soil acidification, increased mobilization of Al from soil
June 2018
5-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
to drainage water, and depletion of the pool of exchangeable base cations in the soil (see
Appendix 4.3.4 and Appendix 4.3.5 for descriptions of these processes). Effects on the
soil and direct effects of acidifying deposition on foliage can influence the response of
plants to climatic stresses such as drought and cold temperature. The effects of acidifying
deposition can also influence the sensitivity of plants to other stresses, including insect
pests and disease (Joslin et al.. 1992).
Since publication of the 2008 ISA, the overarching understanding of terrestrial
acidification has not appreciably changed. More recent research has confirmed and
strengthened this understanding and provided more quantitative information, especially
across the regional landscape. This appendix highlights findings from the literature after
the completion of the 2008 ISA. A number of studies have evaluated the relationships
between soil chemistry indicators of acidification and ecosystem biological endpoints
(Table 5-6). Much of the new literature reviewed in Table 5-6 concerns natural variability
in soil pH, Ca concentrations and studies that have used elevated N and S have been
noted in that table under the "N and S Deposition/Additions" column. Soil chemistry
indicators examined in recent literature include exchangeable base cations, soil pH,
exchangeable acidity (hydrogen ions [H+] and Al), exchangeable Ca:Al ratio, base
saturation, and Al concentrations. Biological endpoints evaluated included physiological
responses of trees and other vegetation, lichens, soil biota, and fauna. Trees and other
vegetation included sugar maple (Acer saccharum), red spruce (Picea rubens), yellow
birch (Betula alleghaniensis), American beech (Fagus grandifolia), American basswood
(Tilia americana), black cherry (Prunus serotina), eastern hophornbeam (Ostrya
virginiana), white ash (Fraxinus americana), hickories (Carya spp.), northern red oak
(Quercus rubra), and forest understory, grassland, and alpine plant species. Table 5-1
provides a summary of the soil chemistry indicator-biological endpoint relationships that
have been evaluated in the literature since the 2008 ISA. The 2008 ISA only considered
changes in vegetation in the analysis of how acid deposition affected terrestrial
ecosystems. More recent research has quantified effects on fauna (e.g., birds, snails) and
soil biota (Appendix 5.2.4; Appendix 5.2.5).
Together with the information available in the 2008 ISA, this body of evidence is
sufficient to infer a causal relationship between acidifying N and S deposition and
the alteration of the physiology and growth of terrestrial organisms and the
productivity of terrestrial ecosystems. Further, the body of evidence is sufficient to
infer a causal relationship between acidifying N and S deposition and the alteration
of species richness, community composition, and biodiversity in terrestrial
ecosystems.
5-2
DRAFT: Do Not Cite or Quote

-------
Transpiration
FUNCTION
Membrane integrity
Stomatal regulation
Enzyme activation
Carbohydrate metabolism
Cold hardiness
Defense/chemical-physical
I
GROWTH
Cell division
Cell wall synthesis
Stress tolerance
I
STRUCTURE
Canopy integrity
Leaf form
Wood quality
Tree height
Root distribution
H'J
Deposition
Cytoplasm j
Cell Wall
Caj
Leaching
Throughfall
, . Root xylem and cortex
Ht'on
s In
x
|Ca. Al

J
Soil
Solution
Ca, Al


Leaching
Physiological Processes
Ca Supply Rate
«	
Biogeochemical Processes
Al = aluminum; Ca = calcium; H = hydrogen; H20 = water; S = sulfur.
Source: Fenn et al. (2006'i.
Figure 5-1 Diagram based on Fenn et al, (2006) showing indicators of forest
physiological function, growth, and structure that are linked to
biogeochemical cycles through processes that control rates of
calcium supply. Calcium affects plant physiological processes
that influence growth rates and the capacity of plants to resist
environmental stresses, such as extremes of temperature,
drought, insects, and diseases. Therefore, acidifying deposition,
which can deplete soil calcium or interfere with calcium uptake
through mobilization of soil aluminum, can affect forest health.
June 2018
5-3
DRAFT: Do Not Cite or Quote

-------
Table 5-1 Relationships between soil chemistry indicators and biological endpoints that have been evaluated
in the literature since the 2008 Integrated Science Assessment.
Taxa
Exchangeable Base
Cations
PH
Exchangeable
Acidity
Exchangeable
Ca:AI Ratio
Base Saturation
Exchangeable Al
or Al
Concentrations
Sugar maple
Beieretal. (2012);
Bilodeau-Gauthier et
al. (2011);
Long et al. (2009);
Page and Mitchell
(2008):
Sullivan et al. (2013);
Cleavitt et al. (2014);
Duchesne and
Ouimet (2009);
Pitel and Yanai
(2014)
Long et al. (2009);
McEathron et al.
(2013); Sullivan et
al. (2013);
Miller and
Watmough (2009)
Bilodeau-Gauthier
et al. (2011)
Bilodeau-Gauthier Bilodeau-Gauthier
et al. (2011);
Sullivan et al.
(2013)
et al. (2011)
Yellow birch
McEathron et al.
(2013)

American beech
Page and Mitchell
(2008);
Duchesne and
Ouimet (2009)

American basswood
Page and Mitchell
(2008);
Beieretal. (2012)
- - - -
Black cherry
-
Long et al. (2009)
Eastern hophornbeam
Beieretal. (2012)
-
Hickories
-
Elias et al. (2009)
June 2018
5-4
DRAFT: Do Not Cite or Quote

-------
Table 5-1 (Continued): Relationships between soil chemistry indicators and biological endpoints that have been
evaluated in the literature since the 2008 Integrated Science Assessment.
Taxa
Exchangeable Base
Cations
PH
Exchangeable
Acidity
Exchangeable
Ca:AI Ratio
Exchangeable Al
or Al
Base Saturation Concentrations
Northern red oak
-
-
-

Elias et al. (2009)
Forest understory plant
Horslev et al. (2008)
Horslev et al. (2008)
Horslev et al. (2008) -

Horslev et al. (2008)
species





Grassland plant species
-
Pannek et al. (2015)
-

-
Lichens
-
Cleavitt et al.
(2011a)
-

-
Soil biota
Sridevi etal. (2012):
Ohta etal. (2014)
Chen etal. (2013):
Rousk et al. (2009):
Gilliam et al.
(2011b):
Sridevi etal. (2012)


Chen etal. (2013):
Rousk et al. (2009)
Fauna
Beier et al. (2012):
Pabian and
Brittinaham (2012)
Pabian and
Brittinaham (2012)


-
Al = aluminum; Ca = calcium.
Note: only soil chemistry indicators with a reported significant relationship (positive or negative) with a biological endpoint are indicated in this table. See Table 5-6 for a listing of soil
chemical indicators included in each study.
June 2018
5-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
5.2.
Effects on Terrestrial Organisms and Ecosystems
5.2.1. Trees and Forests
Both coniferous and deciduous forests throughout the eastern U.S. have been
experiencing gradual losses of base cation nutrients from the soil due to accelerated
leaching from acidifying deposition. This change in base cation nutrient availability can
reduce the quality of forest nutrition over the long term in sensitive areas of low base
cation soils. Evidence suggests that red spruce and sugar maple in some areas in the
eastern U.S. have experienced declining health as a consequence of acidifying deposition.
The 2008 ISA reported several indicators of stress to forest trees (Table 3-3 in the 2008
ISA), including the percentage of dieback of canopy trees, dead tree basal area (as a
percentage), crown vigor index, and fine twig dieback. Biological effects of acidification
on terrestrial ecosystems were generally attributed to A1 toxicity, decreased ability of
plant roots to take up nutrient cations (due to leaching of base cations from soil and
interference with uptake), and elevated leaching of Ca2+ from conifer needles. The Ca:Al
ratio in soil solution is a chemical indicator of the negative impacts of soil acidification
on terrestrial vegetation (Cronan and Grigal. 1995). As tree species have shown similar
sensitivities to Ca:Al and the molar ratio of base cation (Bc):Al in soil solution (with Be
representing Ca2+, Mg2+, and K+), the Be: A1 ratio was used to represent the Ca: A1 ratio,
and it is the most commonly used indicator in the simple mass balance (SMB) model to
estimate critical acid loads in the European Union, Canada, and the U.S. (McNultv et al..
2007; Ouimet et al.. 2006; Spranger et al.. 2004).
Sverdrup and Warfvinge (1993). in a meta-analysis of laboratory, greenhouse, and field
studies, reported that the critical soil solution Be: Al ratios for a large variety of tree
species ranged from 0.09 to 20.0, although the Bc:Al ratio range reported for tree species
native to North America was more restricted at 0.09 to 2.0. This range is similar to that
described by Cronan and Grigal (1995) for Ca:Al. In their meta-analysis of studies
examining sensitivities to the soil solution Ca: Al ratio, plant toxicity or nutrient
antagonism was reported to occur at Ca:Al ratios ranging from 0.2 to 2.5. The
meta-analyses conducted by Sverdrup and Warfvinge (1993) explored the relationships
between Bc:Al ratios in soil solution and the growth of different tree species. They
reported the Bc:Al ratios at which growth was reduced by 20% relative to controls. A
Bc:Al ratio of 1.0 is often applied to protect forested systems of Europe, particularly
conifers (Spranger et al.. 2004). and a Bc:Al ratio of 10.0 has been identified for forests
in North America, to protect deciduous forests (McNultv et al.. 2007; Ouimet et al..
June 2018
5-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
2006). The higher ratio (i.e., 10.0 vs. 1.0) provides a greater level of protection for a
wider range of species and various biotic and abiotic conditions.
However, when using soil solution Bc:Al or Ca:Al ratios as the chemical indicators or
criterion in the estimation of terrestrial acidification critical loads, several uncertainties
must be considered. Many of these considerations have been addressed in the reviews
conducted by Cronan and Grigal (1995) and Vanguclova et al. (2007). Although
recognized as one of the best indicators of soil acidity risk in forest ecosystems, soil
Ca:Al and Bc:Al ratios are variable and can change with soil conditions temporarily and
spatially. Estimates of soil base cations and Al concentrations are also influenced by
sampling and laboratory analysis methods, and there is still uncertainty regarding which
forms of Al are phytotoxic. Soil solution Al occurs in many different ionic and
complexed forms, depending on the soil pH and concentrations of soil ligands. In
addition, uncertainties regarding the sensitivity of the biological receptor should also be
considered. Critical Bc:Al and Ca:Al ratios are often based on seedling studies in
controlled environments, and the relationships are less consistent for trees growing in the
forest. Environmental and biological conditions, such as differences in tree age, soil
horizon chemistry experienced by the roots, and root mycorrhizae are important to
consider when comparing laboratory and field-based research studies. Recognizing these
different sources of variability, Cronan and Grigal (1995) recommended applying a ±50%
uncertainty to their critical soil Ca:Al ratio of 1.0. Critical fine root Ca:Al ratios have also
been suggested as indicators of stress in acidic forest soils (Vanguclova et al.. 2007;
Cronan and Grigal. 1995). Fine root Ca and Al indicate what is absorbed by the tree, and
there are often strong correlations between fine root and soil solution Ca:Al ratios.
Similarly, studies have shown relationships between fine root Ca:Al ratio and metrics of
root and above-ground tree health and productivity. However, similar to critical soil
solution Ca:Al and Bc:Al ratios, fine root Ca:Al ratios also have sources of uncertainty
that can be attributed to factors including soil water and chemistry variability, analysis
methods and timing, fine root age, and differences in tree physiology. In addition, EPA is
unaware of a method that uses fine root Ca: Al ratio as the chemical indicator within SMB
calculations of critical loads.
The tree species most commonly studied in North America and used in the 2008 ISA to
assess the impacts of acidification due to total nitrogen and sulfur deposition include red
spruce and sugar maple, although other tree species like flowering dogwood (Cornus
florida) can also be sensitive to conditions created by acidifying deposition. New
information regarding the effects of acidifying deposition on the commonly studied sugar
maple and red spruce, as well as other tree, understory, and grassland species is
summarized below.
5-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
5.2.1.1. Sugar Maple
Among broadleaf tree species in the northeastern U.S., sugar maple is the species most
commonly associated with adverse acidification-related effects resulting from S and N
deposition, although other base cation accumulating hardwoods may also be at risk
(Driscoll et al.. 2001b). Sugar maple is distributed from Missouri and Minnesota in the
northcentral U.S. eastward to Maine and the central Appalachian Mountain region, and
the species is a dominant component of the northern hardwood forest type (Braun. 2001).
Within this range, soil acidification is widespread throughout the northeastern U.S. and
within many portions of the Appalachian Mountains ("Warbv et al.. 2009).
The 2008 ISA reported that acidifying deposition, in combination with other stressors, is
a likely contributor to the decline of sugar maple trees (Section 3.2.2.3 of 2008 ISA).
Sugar maple decline has been noted to occur in some portions of the eastern U.S., on base
cation-poor soils developed from parent material dominated by sandstone or other base
cation-poor substrates. Sugar maple appears to be particularly sensitive to reduced Ca and
magnesium (Mg) availability caused by acidifying deposition. Based on the results from
a compilation of laboratory studies, sugar maple growth can be reduced by approximately
20% at a Bc:Al soil solution ratio of 0.6 (Sverdrup and WarfVinge. 1993).
The more recent literature on sugar maple is consistent with the 2008 ISA. In these
studies, sugar maple was sensitive to soil conditions and chemistries associated with
acidifying deposition (Appendix 5.2.1.1.1). Soil chemical indicators that were evaluated
included exchangeable base cations, soil pH, exchangeable acidity (H+ and Al),
exchangeable Ca:Al ratios, base saturation, and Al concentrations (Al form unspecified).
Measured sugar maple responses included changes in basal area, growth, regeneration
success, and foliar nutrient concentrations and chemistry (Table 5-6). In addition, several
studies evaluated physiological mechanisms that could explain the response of sugar
maple to changes in soil chemistry induced by acidifying deposition (Table 5-2).
June 2018
5-8
DRAFT: Do Not Cite or Quote

-------
Table 5-2 Summary of calcium addition studies in North America.
Reference
Region
Species
Additions
(Ca and/or Al)
Description
Battles et al.
New
Sugar maple,
Approximately
Ca additions resulted in the recovery of tree biomass increments, higher aboveground NPP,
(2014)
Hampshire
American
1,000 kg Ca/ha
and increased photosynthetic surface area. Sugar maple exhibited the largest cumulative


beech, yellow
(applied in 1999)
change in biomass, while American beech showed a negative cumulative change in


birch, red

biomass.


spruce, and




balsam fir


Bovce et al. New	Red spruce Total of 38 g Ca/m2 Trends toward greater foliar Ca and Ca:AI ratios and lower Al concentrations across the
(2013)	Hampshire, and balsam fir and/or 10.8 g Al/m2 treatment gradient. Ca availability appeared to enhance the ability of red spruce and balsam
Vermont	fir to repair oxidative stress damage.
Cleavitt et
al. (2011b)
New
Hampshire
Sugar maple
Approximately
1,000 kg Ca/ha
(applied in 1999)
Masting events were not influenced by the Ca treatment. Seeds from the Ca treated sites
had lower concentrations of Al, K, and Mg and significantly higher concentrations of Si (but
no differences in Ca concentrations). Seeds from the Ca treated sites had higher
percentage of seedling emergence and higher seedling survival for the first 3 yr.
Halman et
al. (2008)
New
Hampshire
Red spruce
Approximately
1,000 kg Ca/ha
(applied in 1999)
Foliar Ca and total sugar concentrations were significantly higher in trees in the Ca addition
watershed. Cold tolerance of foliage was significantly greater in trees in the Ca addition
watershed.
Halman et
al. (2013)
New
Hampshire
Sugar maple
Total of 38 g Ca/m2
and/or 10.8 g Al/m2
Ca additions were found to increase C allocation to sugar maple growth. Al additions
increased root Al concentrations, root cell membrane disturbance, and ascorbate
peroxidase and glutathione reductase activity, and reduced foliar reflush following frost
injury and the number of viable seeds.
Halman et
al. (2015)
New
Hampshire
Sugar maple
and American
beech
Annual application of
CaCb (2.5 g/m2) and
AlCb (0.9 g/m2) from
1995-1999;
Wollastonite (38 g/m2)
was applied in 1999 a
single dose
The Al and Ca treatments did not affect beech foliar chemistry. Ca additions resulted in
significantly higher Ca concentrations in both dominant and nondominant sugar maple trees.
By 2008, the growth of American beech was higher than that of sugar maple on the control
plots and Al treated plots, and nondominant sugar maple growth was greater than that of
American beech and dominant sugar maple on the Ca treated plots. Increases in tree
mortality on the Al treated plots may have released surviving American beech and increased
their growth.
June 2018
5-9
DRAFT: Do Not Cite or Quote

-------
Table 5-2 (Continued): Summary of calcium addition studies in North America.
Reference
Region
Species
Additions
(Ca and/or Al)
Description
Minocha et
al. (2010)
New
Hampshire
Sugar maple,
yellow birch,
and American
beech
Approximately
1,000 kg Ca/ha
(applied in 1999)
Foliar soluble Ca increased significantly in all species at mid and high elevations at Ca
supplemented watershed. Sensitivity to Ca limitation increased with elevation.
Moore et al.
(2012)
Quebec,
Canada
Sugar maple
8 dolomite lime
addition rates—0, 0.5,
1, 2, 5, 10, 20, and
50 megatonnes/ha
(applied in 1994)
Foliar Ca and Mg concentrations were found to be higher for treated trees relative to the
control trees. Mean crown dieback of trees decreased and seedling density increased with
the lime rate. The stem basal area increment for maple trees and proportion of the sugar
maple seedlings was increased with lime treatment.
Schabera et
al. (2011)
New
Hampshire
Red spruce
Approximately
1,000 kg Ca/ha
(applied in 1999)
Trees from the Ca addition watershed had higher estimated levels of foliar biomass. Ca
addition increased the stress tolerance of red spruce foliage during the cold season,
resulting in greater crown mass.
Smith et al.
(2009)
New York,
Vermont
Red spruce
Total of 160 kg Ca/yr
(1992-1995)
Greater amounts of Ca were found in the wood from the high-Ca location than from the
low-Ca location. Ca concentration was greater in wood formed in the 1970s than for wood
formed in the 1980s, and Ca treatments resulted in increased concentration of Ca in both
the 1970s and 1980s decadal bands of wood. Foliar concentrations ofCa oxalate were also
higher on the high-Ca site.
Sridevi et al.
(2012)
New
Hampshire
Soil microbes
(bacteria)
Approximately
1,000 kg Ca/ha
(applied in 1999)
Bacterial community structure in the Ca treated and nontreated reference soils was found to
be significantly different, with differences among communities being more pronounced in the
mineral soils.
Al = aluminum; AICI3 = aluminum chloride; C = carbon; Ca = calcium; CaCI2 = calcium chloride; g = gram; ha = hectare; K = potassium; kg = kilograms; m = meter; Mg = magnesium;
NPP = net primary production; Si = silicon; yr = year.
June 2018
5-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Soil Chemical Indicators for Sugar Maple
Soil Exchangeable Base Cations
Sugar maple basal area and growth were responsive to soil exchangeable base cation
concentrations in several studies in the northeastern U.S. In the Adirondack Mountains of
New York, Beier et al. (2012) reported sugar maple basal area was positively correlated
to forest floor and mineral soil (B-horizon) exchangeable soil Ca on sites that ranged in
soil Ca from 1.83 to 53.89 cmolc/kg (in Oa-horizon) and 0.28 to 7.73 cmolc/kg in the
B-horizon (see Figure 3-2 in the 2008 ISA for a description of soil horizons). Similar
results were reported in Quebec, Canada by Bilodeau-Gauthier et al. (2011) who found
that sugar maple basal area growth was positively correlated to concentrations of base
cations (Ca, potassium [K], and Mg) in wood and mineral soil (B-horizon), and Long et
al. (2009) who reported that sugar maple basal area increment (BAI) was positively
correlated with exchangeable Ca and Mg in the upper B-horizon in the northeastern U.S.
Page and Mitchell (2008) also found that the relative basal area of sugar maple was
positively correlated with mineral soil (0-10 cm) exchangeable Ca in the Adirondack
Mountains. In the same region, Sullivan et al. (2013) determined that sugar maple canopy
vigor was positively correlated with soil exchangeable Ca and Mg, and mean growth
rates (measured as BAI) were positively correlated with exchangeable Ca at the
watershed level.
Relationships between soil exchangeable base cations and sugar maple regeneration (the
growth and abundance of seedlings and saplings) have also been demonstrated. Cleavitt
et al. (2014) evaluated sugar maple seedling survival and cause of death across 22 sites in
New Hampshire that varied in soil Ca and topographic position. Soil Ca concentration
exhibited a nine times change across the study sites and was positively correlated to
increases in sugar maple abundance and initial seedling densities. However, soil Ca
concentration was not a significant predictor of first-year mortality, nor was it a factor
that distinguished among the three main site types. In a study in Quebec, Canada,
Duchesne and Ouimet (2009) explored relationships between soil chemistry and sugar
maple in the sapling stratum and found that the basal area of sugar maple was positively
correlated with soil exchangeable Ca and Mg.
Pitel and Yanai (2014) evaluated the abiotic and biotic factors influencing the mortality
of dominant and codominant sugar maple trees in 47 stands in Massachusetts, Vermont,
and New York that had experienced defoliation by native forest tent caterpillars
(Malacosoma disstria) between 2002 and 2007. A total of 54 predictive variables
(defoliation year, stand and site characteristics, and soil chemistry variables) were
5-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
evaluated through multiple linear regression followed by stepwise regression to produce
the best predictive models of sugar maple mortality from 2006 to 2008. Mortality was
highest in stands with the greatest amount of crown dieback the previous year, and
drought, cold winter temperatures, concave micro-relief, and soil base cation availability
were significant predictors of mortality. Concentrations of exchangeable Ca, Mg, and K
in the upper B soil horizon were inversely correlated with sugar maple mortality, with
exchangeable K showing the strongest relationship. Sites with above average sugar maple
mortality (>3 or 4%) were found to occur on soils with low concentrations of
exchangeable Ca (0.31-0.46 cmolc/kg), Mg (0.06-0.10 cmolc/kg), and K
(0.03-0.05 cmolc/kg). There was also an interaction between defoliation and soil base
cation availability: stands defoliated in 2005 that had low soil (A-horizon) Mg saturation
suffered higher rates of mortality.
Soil Exchangeable Acidity and pH
Bilodeau-Gauthier et al. (2011) found that sugar maple tree growth was negatively
correlated to soil exchangeable acidity (H+ and exchangeable Al), and through a
multifactor analysis, showed that tree age and soil exchangeable Al accounted for 51% of
the variation in sugar maple BAI. Positive correlations between mineral soil pH and sugar
maple basal area growth have also been reported by Long et al. (2009) and McEathron et
al. (2013). Sullivan et al. (2013) found that sugar maple canopy vigor was positively
correlated with soil pH.
One study reported a relationship between sugar maple foliar chemistry and
exchangeable soil chemistry. Miller and Watmough (2009). in their evaluation of
hardwood plots along air-pollution (N, S, and ozone), soil-acidity, and climate gradients
in Ontario, Canada, showed that foliar Ca and Mg content was positively correlated and
foliar manganese (Mn) content was negatively correlated with soil A-horizon pH.
Soil Base Saturation
An investigation of every relationship between sugar maple BAI and soil variables
revealed that base saturation was the best predictor of BAI (nonlinear) and explained
43% of variance (Bilodeau-Gauthier et al.. 2011). Similarly, Sullivan et al. (2013) found
that soil base saturation was related to sugar maple regeneration and growth. Plots with
lower soil base saturation did not have sugar maple regeneration; the proportion of sugar
maple seedlings dropped substantially at base saturation levels less than 20%
(Figure 5-2). Mean growth rates were positively correlated with soil base saturation at the
watershed level.
5-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
c
o
r
o
Q.
O
O)
C
0
a>
CO
(/)
0.8 -
0.6 -
0.4
0.2 -
0
0
wi
CO I
o«
CMl
COl
CQ,
%
'<9
5
o
20
40
60
80
100
BS in Upper B Horizon
BS = base saturation; SM = sugar maple.
Note: plots were rank ordered based on soil base saturation, and a 5-plot rolling average was applied to both the soil base
saturation and the seedling proportion data. Reference lines are added at base saturation values of 12 and 20%, indicating break
points in the sugar maple seedling proportion-response function.
Source: Sullivan et al. (2013).
Figure 5-2 Relationship between the proportion of seedlings that were sugar
maple and soil base saturation in the upper B-horizon.
Soil Exchangeable Calcium:Aluminum Ratio
The relationship between sugar maple BAI and a soil exchangeable Ca:Al threshold of
<0.03 was evaluated by Long et al. (2009) through an analysis of a regional network of
76 forest sites that spanned Pennsylvania, New Hampshire, New York, and Vermont.
However, a repeated-measures analysis did not reveal statistically significant effects of
the Ca:Al soil ratio threshold of <0.03 on BAI.
5.2.1.1.2.	Physiological Mechanisms for Sugar Maple
A number of studies were also conducted to determine the potential mechanisms
underlying the sensitivity of sugar maple to the soil conditions created by acidifying
deposition. These studies evaluated the influence of Ca additions on sugar maple
physiology, growth, seeds, and seedlings (Table 5-2). In Hubbard Brook Experimental
Forest (HBEF), NH, one study involved the addition of a total of approximately 1,000 kg
Ca/ha as pelletized wollastonite (CaSiCh) in October 1999 to a watershed (Battles et al..
June 2018
5-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
2014). This treatment resulted in increases in: soil pH (in Oie-horizon) from 4.29 to 5.45;
base saturation from 9.6 to 78.6% (in Oie layer), 56% (in Oa-horizon), and 14.3% (in
upper mineral soil); and soil solution Ca:Al (inorganic monomeric Al) from 1.6 to 6.4 (in
Oa-horizon) and from 0.97 to 3.8 (in the mineral soil). As part of this study, Battles et al.
(2014) evaluated the effects of the Ca additions on tree growth and found that sugar
maple in the treated watershed exhibited the largest increase in cumulative biomass
relative to other species. Another study at the same experiment reported that sugar maple
seedlings were 50% larger in the treated watershed and mycorrhizal colonization of
seedlings was much higher in the treated watershed (22.47% of root length) as compared
with the reference watershed [4.4%; (Juice et al.. 2006)1. Mycorrhizal colonization also
increased in mature sugar maple in the treated watershed (56%) compared to the
reference watershed [35%; (Juice et al.. 2006)1. In contrast to sugar maple, American
beech showed a negative cumulative change in biomass. Minocha et al. (2010) examined
the effects of the Ca addition treatment on foliar chemistry and found that sugar maple
had increased foliar concentrations of Ca, which were accompanied by increases in
soluble phosphorous (P), chlorophyll, and two amino acids (i.e., glutamate and glycine).
The authors also reported decreases in known metabolic indicators of physiological stress
(i.e., arginine and y-aminobutyric acid, as well as putrescine, a diamine). Sugar maple
also exhibited a decrease in foliar putrescine at higher elevations in the watershed,
indicating possible remediation from Ca deficiency. Cleavitt et al. (201 lb) evaluated
sugar maple seed production, seed chemistry, seedling growth (lifestage), and seedling
survival on the Ca treated watershed and found that seeds from the Ca treated sites had
lower concentrations of Al, K, and Mg and significantly higher concentrations of silicon
(Si) than seeds from the reference watershed. The percentage of seedling emergence was
also higher, average lifestage was more significantly advanced, and seedling survival was
greater on the Ca treated watershed.
In a second study (NuPert Study) conducted at HBEF, Halman et al. (2013) evaluated the
impacts of Ca (380 kg/ha) and Al (108 kg/ha) additions on sugar maple physiology. The
Ca additions increased the proportion of carbon allocated to sugar maple growth. In
contrast, Al additions increased root Al concentrations and root cell membrane
disturbance, increased ascorbate peroxidase (APX) and glutathione reductase (GR)
activity, reduced foliar reflush following frost injury, and reduced the number of viable
seeds. Ascorbate peroxidase and GR are antioxidant enzymes that target reactive oxygen
species generated by environmental stresses, such as photo-oxidative stress damage at
low temperatures. Elevated APX and GR activity are indicators of increased oxidative
stress. Several years later, Halman et al. (2015) re-evaluated the NuPert study to
determine the long-term and contrasting responses of sugar maple and American beech to
the Ca and Al treatments following a major ice storm in 1998. The Al and Ca treatments
did not affect beech foliar chemistry. However, Ca additions significantly increased Ca
5-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
concentrations in both dominant and nondominant sugar maple trees, relative to the trees
treated with Al. Interestingly, for all treatments, foliar A1 concentrations were higher in
the nondominant than dominant sugar maple trees, and the opposite was found for foliar
Ca concentrations. By 2008 (11 years after the ice storm), the growth of American beech
was higher than that of sugar maple on the control plots and Al treated plots, and
nondominant sugar maple growth was greater than that of American beech and dominant
sugar maple on the Ca treated plots. These differential growth responses emerged within
2 to 11 years following the ice storm, depending on the treatment and species. Although
plots were mainly composed of sugar maple, American beech experienced the greatest
growth on Al treated plots. Increases in overstory tree mortality on the Al treated plots
may have increased light availability, released surviving American beech, and increased
beech growth.
Moore et al. (2012) evaluated soil chemistry and sugar maple status 15 years after
treatment with dolomite lime (0 to 50 megatonnes/ha) to a hardwood forest in Quebec,
Canada, and found results similar to those in HBEF. Foliar Ca and Mg concentrations
were higher for treated trees relative to the control trees, and mean crown dieback of trees
decreased quadratically with the lime addition rate, from 39% for the control trees to a
value of 1 to 3% for the lime rates of 5 megatonnes/ha and higher. In addition, sugar
maple BAI for the limed trees was nearly double that of the nonlimed trees. The lime
application was also beneficial to the sugar maple regeneration. The overall sugar maple
seedling density increased with the lime rate, doubling in the 50 megatonnes/ha
(32 seedlings/m2) compared with the controls (16 seedlings/m2). The proportion of the
sugar maple seedlings to all of the other species increased quadratically from 22% in
controls to more than 55% in the 5 to 50 megatonnes/ha treatments. In contrast, the
proportion of competitive species decreased quadratically with the lime rate, including
American beech, for which the proportion in the treated plots (24%) was nearly half the
proportion observed in the controls (46%).
5.2.1.2. Red Spruce
Red spruce is a conifer that occurs mainly in the northeastern U.S. and at scattered
high-elevation sites (e.g., mountain and ridge tops) in the Appalachian Mountains. Red
spruce dieback or decline has been observed across high elevation landscapes of the
northeastern, and to a lesser extent, southeastern U.S. At high elevations in the
Adirondack Mountains in New York and the Green Mountains in Vermont, more than
50% of the canopy red spruce trees died during the 1970s and 1980s. In the White
Mountains in New Hampshire, about 25% of the canopy spruce died during that same
period (Dc Haves etal.. 1999). Dieback of red spruce has also been observed in mixed
June 2018
5-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
hardwood-conifer stands at relatively low elevations in the western Adirondack
Mountains, an area that receives high inputs of acidifying deposition (Shortlc ct al..
1997); acidifying deposition has been implicated as a causal factor (DcHaves ct al..
1999).
The 2008 ISA reported that changes in soil chemistry (e.g., depletion of soil base cations,
Al toxicity to tree roots, leaching of base cations into drainage water; see Appendix 4.3.4
and Appendix 4.3.5) have contributed to high mortality rates and decreased growth of red
spruce trees in some areas of the eastern U.S. over the past three decades. Studies
evaluating the physiologic basis behind the responses of red spruce to acidifying
deposition have attributed the reduced vigor of the species to increased sensitivity to frost
injury and cold temperatures. The frequency of freezing injury to red spruce needles had
increased during a period in the latter half of the 20th century that coincided with
increased emissions of S and N oxides and increased acidifying deposition (DcHaves et
al.. 1999).
Since the 2008 ISA, research evaluating the sensitivity of red spruce to soil indicators of
acidifying deposition have mainly focused on the potential mechanisms underlying the
sensitivity of the species to acidic soil conditions. These studies have examined the
physiological response of red spruce to Ca additions (Table 5-2).
In the HBEF study that involved the addition of approximately 1,000 kg Ca/ha, Halman
et al. (2008) and Schaberg et al. (2011) evaluated the response of red spruce foliar
chemistry. Halman et al. (2008) reported significantly higher foliar Ca and total sugar
(fructose and glucose) in the Ca addition watershed than in trees in the reference
watershed. Foliar APX activity was also higher in trees in the Ca addition watershed
during winter. Cold tolerance of foliage was significantly greater in trees in the Ca
addition watershed than in trees in the reference watershed. Schaberg et al. (2011)
measured concentrations of foliar polyamines and free amino acids, foliar chlorophyll,
and sapwood area (as a proxy for foliar biomass) in Ca treated and nontreated trees.
Foliar polyamines (putrescine and spermidine) and free amino acids (alanine,
aminobutyric acid [GABA]) are putative stress protection compounds that may directly
protect or provide other benefits to foliage that increase stress tolerance. The Ca additions
increased November concentrations of alanine, GABA, putrescine, and spermidine
relative to foliage from the reference watershed. Consistent with increased stress
protection indicated by the elevated polyamine and free amino acids, foliage from the Ca
addition watershed had higher total chlorophyll concentrations than foliage from the
reference watershed. In contrast, foliage from the reference watershed had significantly
higher alanine:glutamic acid ratios, which have been attributed to cold sensitivity or
damage in other species. In addition to concentration-based differences in foliar
5-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
compounds, trees from the Ca addition watershed had higher estimated levels of foliar
biomass than trees from the reference watershed. These findings suggested that Ca
addition increased the stress tolerance and productive capacity of red spruce foliage
during the cold season and resulted in greater crown mass compared to trees growing on
nontreated soils.
Similar findings were reported by Bovce etal. (2013). who examined the influence of Ca
and A1 on the physiology of red spruce and balsam fir (Abies balsamea) in three different
locations that varied in soil nutritional status. Processes known to be Ca sensitive (root
and foliar cation concentrations, chlorophyll fluorescence, soluble sugar concentrations,
and the activities of APX and GR in current-year foliage) were measured, and the results
from the study suggest that Ca availability enhanced the ability of red spruce and balsam
fir to repair oxidative stress damage, including photo-oxidation.
Concentrations of Ca in sapwood and accumulations of Ca oxalate in foliage have been
used as markers of environmental change due to acidic deposition or forest management
practices. Smith et al. (2009) compared the effects of different Ca fertilization treatments
(approximately 178 kg Ca/ha/yr during 1992-1995) on Ca concentrations in wood and
Ca and oxalate (Ox) concentrations in red spruce foliage at two locations with different
initial concentrations of Ca in the soil (6.4 cmolc/kg vs. 13.7 cmolc/kg). Greater amounts
of Ca were found in the wood from the high-Ca location than from the low-Ca location.
Foliar concentrations of Ca oxalate were higher on the high-Ca site than on the low-Ca
site and increased in response to Ca additions.
5.2.1.3. Other Tree Species
In the 2008 ISA, there was some information regarding the effects of acidification on
dogwood (Cornus) species. Loss of base cations, specifically Ca2+, had been implicated
in increased susceptibility of flowering dogwood to dogwood anthracnose, its most
destructive disease. Susceptibility to the disease and disease severity in stands appeared
dependent on several factors, including acid deposition and various edaphic
characteristics and meteorological conditions. Studies pointed to greater vulnerability of
dogwoods to anthracnose under simulated acid rain treatments (Britton et al.. 1996) and
under Ca2+ deficiency KHolzmucllcr et al.. 2007); Section 3.2.2.3 of 2008 ISA],
Since the 2008 ISA, no further research on dogwood has been published. However, other
tree species have been evaluated in studies relating soil chemistry to tree physiology,
including red maple, white oak, yellow birch, white ash, American beech, black cherry,
northern red oak, hickories, American basswood, and eastern hophornbeam. These
June 2018
5-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
species were reported to vary in their sensitivities to soil conditions associated with
acidifying deposition (Table 5-6).
In a recent study in West Virginia, Thomas et al. (2013) used dendroisotopic techniques
to show the recovery of eastern redcedar (Juniperus virginiana) trees from decades of S
pollution using an analysis of a tree ring chronology from 1909 to 2008. Growth,
measured as BAI, in eastern redcedar old-growth stands (118-480 years old) increased
significantly since 1970. A multivariate correlation analysis using historical climate
variables, atmospheric CO2 concentrations, and U.S. SO2 and NOx emissions estimates
showed that the growth of cedar trees over the 100-year chronology is explained best by
increases in atmospheric CO2 and NOx emissions, and decreases in SO2 emissions.
Through carbon isotope (13C) analyses, the researchers were able to show that the stomata
of cedar may be more sensitive to SO2 emissions than to increasing atmospheric CO2. A
breakpoint in the 100-year S13C tree ring chronology occurred around 1980, as SO2
emissions declined, indicating a gradual increase in stomatal conductance and a
concurrent increase in photosynthesis related to decreasing SO2 emissions and increasing
atmospheric CO2. After 1990, calculated stomatal conductance increased more than
photosynthesis. These patterns in physiology led to changing trends in intrinsic water use
efficiency (/WUE; the ratio of photosynthesis to water loss through stomatal
conductance). The calculated /'WUE generally increased from the 1940s to 1990 mainly
because of SO2 emission effects on stomatal conductance, and then /'WUE began to
decrease after 1990 because SO2 emissions no longer constrained stomatal function.
Through S isotope analysis (tree ring S34S), the study showed a synchronous change in
the sources of S used at the whole-tree level that indicated a reduced anthropogenic
influence. The increase in growth and the S13C and S34S trends in the tree ring chronology
of these Juniperus trees provide evidence for a distinct physiological response to changes
in atmospheric SO2 emissions since 1980. The authors attributed the changes since 1980
to an indirect effect of decreases in acid deposition. However, the exact mechanism is
unclear because the 100-year chronology could only be correlated to estimated SO2
emissions because acidifying deposition measurements were not available. Other
researchers have pointed out that the trees in the (Thomas et al.. 2013) study were
growing on a limestone outcrop that could be well buffered from soil acidification
(Schaberg et al.. 2014). Further, the rapid recovery of tree growth could also point to a
direct effect of gaseous SO2 rather than an indirect effect of soil acidification, which
would have a longer response time to decreases in emissions. See Appendix 3_2 for
further discussion of direct gaseous SO2 effects on vegetation.
5-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Soil Chemical Indicators for Other Tree Species
Soil Exchangeable Base Cations
Several studies reported on the relationships between species-specific basal area and
growth and soil exchangeable base cations. In the Adirondack Mountains, McEathron et
al. (2013) evaluated the relationships between basal area and soil chemistry and reported
that yellow birch basal area was positively correlated with mineral soil exchangeable Ca.
Similarly, Page and Mitchell (2008) compared the basal areas of American basswood,
American beech, and white ash with mineral soil (0-10 cm in depth) exchangeable Ca
concentrations and found exchangeable Ca had a positive correlation with relative basal
area of American basswood, a negative correlation with the relative basal area of
American beech, and no correlation with relative basal area of white ash. The positive
relationship between American basswood and soil exchangeable Ca is consistent with
other studies. Beier et al. (2012). in their evaluation of vegetation communities across
northern hardwood sites that ranged in exchangeable Ca concentrations from 1.83 to
53.89 cmolc/ha (in Oa-horizon) and 0.28 to 7.73 cmolc/ha (in B-horizon), reported that
American basswood and eastern hophornbeam were only found on the sites with the
highest exchangeable Ca concentrations. Similarly, negative relationships between
American beech and soil exchangeable Ca have been reported elsewhere. Duchesne and
Ouimet (2009) found that basal area of American beech in the sapling stratum in sugar
maple-dominated forests was negatively correlated with exchangeable Ca and Mg. Soil
concentrations of Ca and Mg on sites colonized by American beech were, on average,
roughly half those on sites where American beech was absent, thereby suggesting that
beech effectively colonizes sites with low base status. These findings are also consistent
with the responses of American beech to the Ca addition studies described earlier (Battles
et al.. 2014; Moore et al.. 2012).
Soil Base Saturation
One study of 30 forest plots within Monongahela National Forest, WV reported on the
relationship between soil base saturation and tree abundance. Elias et al. (2009) evaluated
relationships between tree growth parameters and soil indicators of acidifying deposition
and found that hickories were the only species (out of four within-species comparisons)
to have significantly lower numbers on sites with base saturation below 20% (A-horizon)
and 2.5% (B-horizon). They also found that the percentage of dead northern red oak was
highest on sites with Al concentrations (A-horizon) above 43 cmolc/kg. The authors
reported that sites in the Year 2000 with subsurface base saturation above 10% had more
unique species.
5-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
Soil Exchangeable Calcium:Aluminum Ratio
A study of 76 hardwood stands across the northeastern U.S. noted the relationship
between exchangeable Ca:Al ratios and BAI. Long et al. (2009) found that, unlike sugar
maple, black cherry BAI was greater in stands with exchangeable Ca:Al ratios
(B-horizon) below the <0.03 threshold adopted by the study.
5.2.2. Forest Understory and Grassland Species
The 2008 ISA did not specifically assess the effect of acidification on forest understory
species due to the lack of studies. Also, the 2008 ISA reported that grasslands were
thought to be less sensitive to acidification than woodlands (Kochv and Wilson. 2001;
Blake et al.. 1999). and grasslands with calcareous soils will be less sensitive than those
with acidic soils (Bobbink et al.. 1998).
Since the 2008 ISA, several studies have evaluated the relationships between soil
chemistry indicators of acidification and forest understory and grassland species
(Table 5-6).
5.2.2.1. Soil Chemical Indicators for Forest Understory and Grassland Species
Soil pH, Exchangeable Base Cations, and Exchangeable Acidity
One study examined the relationship between soil chemistry indicators of acidification
and forest understory species. Horslev et al. (2008) evaluated 35 soil chemistry, stand,
and climatic variables as predictors of understory plant species composition in northern
hardwood stands, and found that a base cation-acid cation nutrient gradient accounted for
71.9% (in New Hampshire and Vermont) and 63% (in Pennsylvania and New York) of
the variation in the nonmetric multidimensional scaling ordination analyses of plant
community composition. Soil Ca, Mg, and pH formed the base end and Al, Mn, K, soil
acidity, and organic matter represented the acid end of the gradient. Based on results from
McNemars' exact test, a total of 50 of the 234 understory species were associated with
the base end of the base cation-acid cation nutrient gradient. These species have value as
indicators of sites at the high end of the base cation nutrient gradient in northern
hardwoods, sites that would be suitable for acid-sensitive species such as sugar maple.
In the U.K., Stevens et al. (2010b) used data from a national survey to evaluate the
species richness of 68 grasslands along an N deposition gradient. Ellenberg R
(reaction-soil pH) and N (soil nutrient) scores and an index of soil acidity preference
June 2018
5-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
were used as the metrics to characterize the plant diversity responses to the soil chemistry
conditions created by N deposition. Although the study did not find any significant
relationships between the Ellenberg and soil acidity index values and N deposition, there
was evidence that soil acidification was contributing to changes in species diversity and
community composition. Soil acidification may have led to decreased nutrient
availability and increased A1 solubility, thus, masking the effects of increased soil N
availability (see also Appendix 6.3.5 for N discussion).
A similar analysis along an N deposition gradient (2.4 to 43.5 kg N/ha/yr) was conducted
by Pannek et al. (2015) using a vegetation data set from 153 seminatural acidic
grasslands in northwestern Europe. Species frequency in response to N and other factors
including soil (0-10 cm) P, pH, NH4+, and NOs and geographical, climatic, and
management factors were evaluated. A second set of data from acidic grasslands in
Germany (392 plots) and the Netherlands (144 plots) containing plots from different time
periods were also included in the analyses to determine whether the results of the spatial
gradient approach coincided with temporal changes in the abundance of species. Out of
44 species included in the study, 16 were found to be affected by N deposition, with 12 of
them exhibiting a decreased abundance response. Increasing soil pH and P influenced 24
and 14 species, respectively, predominantly increasing abundance. Change of species
over time was unrelated to their responses to pH. However, species were found to
significantly decline over time in both Germany and the Netherlands in response to N
deposition, soil P, and NO3 . These results show the influence of N deposition in a
decline for many plant species in seminatural acidic grasslands, although impacts of N
deposition on pH did not appear to be the mechanism of plant responses (see also
Appendix 6.2.5 for N discussion).
5.2.3. Lichens
In the 2008 ISA it appeared that lichen populations were affected in areas with acidifying
deposition (Davies et al.. 2007). However, it was not clear whether effects were due to
direct effects of SO2, N effects, or acidifying deposition. See Appendix 6.3.7 for a
discussion of N deposition on lichens.
We have not identified any new studies published since the 2008 ISA that evaluated the
relationships between conditions created by acidifying deposition and lichen physiology.
However, in Acadia National Park, Cleavitt et al. (201 la) evaluated the interactions
among N and S deposition, tree type, and epiphytic lichen and bryophyte diversity,
biomass, and abundance to (1) document any differences in the depositional
environments (throughfall and cloud water) of red spruce and red maple (Acer rubrum)
June 2018
5-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
trees, (2) relate deposition differences to both chemistry of the bark and corresponding
epiphyte biomass and composition on the bark, and (3) describe any species-specific
relationships that emerged between the epiphytes and bark chemistry (Table 5-6).
Throughfall N and S deposition were higher under spruce trees than under maple trees,
and average S concentrations in cloud water were negatively correlated with bark Ca and
bark Mg. Epiphytic lichen richness was higher on maple trees than spruce trees, and tree
species differed in the number of rare epiphytic species. Several pollution-sensitive
epiphytes were restricted to the maples: the cyanolichens Leptogium cyanescens
(bipartite) and Lobariapulmonaria (tripartite), the large ruffle lichen Parmotrema
crinitum, and the moss Zygodon viridissimus var. rupestris. Cyanolichens only occurred
on maple bark and did not occur on any bark with a pH below 5.02. Apparent overlap
between the bark chemistry of spruce and maple, particularly for samples from higher on
maple boles, suggest a reduction in the area of chemically suitable substratum for
epiphytes.
5.2.4. Soil Biota
Soil biota were not specifically addressed in the 2008 ISA. Since the 2008 ISA, several
studies have evaluated the relationships between soil chemistry indicators of acidification
and soil biota (Table 5-6). The effects of elevated N on soil biota is addressed throughout
Appendix 6 (see Appendix 6.2.3. Appendix 6.2.4. Appendix 6.2.5. and Appendix 6.3.5).
5.2.4.1. Soil Chemical Indicators for Soil Biota
Soil pH and Exchangeable Calcium:Aluminum Ratio
Bardhan et al. (2012) evaluated the relationships between soil chemistry and bacterial
diversity in 30 plots along a soil chemistry and deposition gradient in high-elevation
spruce-fir forests in the Great Smoky Mountain National Park (GSMNP). Modeled S
deposition ranged from 6 to 41 kg S/ha/yr, measured soil pH ranged from 3.0 to 4.6, and
measured CEC ranged from 1.3 to 23.1 cmolc/kg (soil samples pooled from the 0-, A-,
and B-horizons). However, bacterial diversity and community composition did not
change along the gradients of S deposition, soil pH, or exchangeable Ca:Al ratio.
Soil pH and Aluminum Concentrations
Chen et al. (2013) evaluated the impacts of S additions on soil microbial and nematode
communities in a grassland system of Inner Mongolia. Seven treatments of 28 to 166 mol
June 2018
5-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
H+/ha were applied in three doses in the form of sulfuric acid (H2SO4) from 2009 to 2010.
The authors found that the proportions of soil bacteria, fungi, and nematodes were altered
by the treatments, and the responses were related to soil pH and A1 concentrations.
Increases in fungal fatty acids (49% increase) and fungi:bacteria ratio (120% increase)
and decreases in total and bacterial fatty acids (40-47% decrease) relative to the controls
were attributed to soil pH and A1 concentrations. High A1 concentrations (51 to
83 mg/kg) were associated with decreases in total fatty acids and bacterial fatty acids and
increases in fungal fatty acids. These results are consistent with a study of an agricultural
soil gradient that showed that at soil pH below 4.5, microbes (bacteria and fungi) are
decreased due to increased A1 solubility (Rousk et al.. 2009). and that fungi are often
more dominant in acidic soils. Chen et al. (2013) also reported that acid additions also
impacted the soil nematode community by initially increasing the total number of soil
nematodes and then altering nematode community composition: bacterivorous and
fungivorous nematodes increased, but plant-feeding and omnivorous and carnivorous
nematodes decreased. The shifts in the nematode community were attributed to decreased
soil pH and changes in soil moisture.
Soil pH and Exchangeable Calcium
One study also investigated the influence of Ca addition (-1,000 kg Ca/ha/yr in 1999) on
the soil bacterial community in a northeastern hardwood forest (Sridevi et al.. 2012). The
study detected 1,756 taxa spanning 42 phyla, 53 classes, 127 orders, and 154 families in
the Ca treated and reference watersheds. Bacterial community structure was significantly
different between the Ca treated and nontreated reference soils, with differences among
communities being more pronounced in the mineral soils. Calcium additions resulted in
significant changes in bacterial community composition in the organic and in the mineral
soil horizons. The numbers of detectable taxa in families such as Acidobacteriaceae,
Comamonadaceae, and Pseudomonadaceae were lower in the Ca amended soils, while
Flavobacteriaceae and Geobacteraceae were higher. Analyses of relationships between
soil chemistry and the bacterial communities indicated that only exchangeable Ca, pH,
and P were significantly correlated with bacterial community structure.
Soil pH
Gilliam et al. (201 lb) evaluated the microbial community across a soil weathering
gradient in a northern hardwood forest in West Virginia. Microbial community
composition, characterized through phospholipid fatty acid (PLFA) analysis, varied
among sites. Fungi were dominant at the most weathered, low NO3 -production site,
while Gram-negative bacteria were significantly higher at the less weathered, moderate
5-23
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
and high NOs -production sites. Accordingly, the fungi:bacteria ratio increased in the
direction of the low NOs -production plots in ordination space. Correlations between the
soil parameters and PLFA results suggest that low soil pH and NO, concentrations
supported fungal dominance, although other important factors including differences in
plant community and clay and organic matter content may have also influenced the soil
microbial community.
Soil Exchangeable Calcium
A study in Japan evaluated the relationship between forest type and soil invertebrate
communities in evergreen broad-leaved forests versus Japanese cedar (Cryptomeria
japonica) plantation forests (Ohtaetal.. 2014). Exchangeable Ca was found to be
significantly higher in soil from the Japanese cedar plantations than the evergreen forest.
The invertebrate community composition also differed significantly between the two
forest types and was best explained by exchangeable Ca concentrations. Two major taxa
of soil crustaceans, Talitridae and Ligidium japonicum, were only found in the Japanese
cedar plantations. In contrast, millipedes (Paradoxosomatidae) and beetles were relatively
abundant in the evergreen plots.
5.2.5. Fauna
Fauna were not specifically assessed in the 2008 ISA. Since the 2008 ISA, two studies
were found that evaluated the relationships between soil chemistry indicators of
acidifying deposition and fauna (Table 5-6).
5.2.5.1. Soil Chemical Indicators for Fauna
Soil Exchangeable Calcium
Beier etal. (2012) characterized the variation in gastropod, salamander, and vegetation
communities among northern hardwood forests attributable to soil exchangeable Ca. The
sites represented the variability in soil Ca in the Adirondack Mountains, ranging from
1.83 to 53.89 cmolc/kg (Oa-horizon) and 0.28 to 7.73 cmolc/kg (B-horizon). Snail
community richness and the abundance and live biomass of red-backed salamanders
(Plethodon cinereus) were all positively correlated with soil Ca. Land snail species
richness and abundance were positively correlated with Oa-horizon Ca and negatively
correlated to SO42 deposition and site elevation (and NO, deposition for snail
abundance). Salamander communities were dominated by mountain dusky salamanders
June 2018
5-24
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
(Desmognathus ochrophaeus) at Ca poor sites, with composition continuously shifting
toward dominance by red-backed salamanders as Ca availability increased. Several
known calciphilic species of snails (Paravitrea multidentata, Gastrocopta pentodon, and
Euconulus polygyrus) were found only at the highest Ca sites. Some of the observations
(e.g., decreasing snail abundance) were confounded by a strong positive correlation
between elevation and estimated acid deposition. However, although the underlying
mechanisms require further study, these findings indicate that Ca availability, which is
shaped by geology and acidic deposition inputs, influences northern hardwood forest
ecosystems at multiple trophic levels.
Soil Exchangeable Calcium and pH
Pabian and Brittingham (2012) determined the relationships between soil (Oa-horizon)
chemistry and forest bird community composition, abundance, and diversity, and
evaluated potential mechanisms responsible for the relationships in oak and red
maple-dominated forests in Pennsylvania. Mean soil (Oa-horizon) exchangeable Ca for
the 14 forest sites ranged from 5.28 to 23.5 meq/100 g and pH ranged from 3.6 to 5.1.
Bird community composition (species richness and species abundances) varied with soil
Ca and pH, with 10 bird species having the highest abundances in forests with high-Ca
soils, and 5 species having the highest abundances with low-Ca soils. Five species were
classified as "generalists" because they had high abundances and were present at all
forest sites. Bird species associated with low-Ca soils were associated with high densities
of mountain laurel (Kalmia latifolia) and five tree species whose basal areas were
explained by low soil pH and Ca. Bird species associated with high-Ca soils were
associated with high densities of saplings and high basal areas of acid-sensitive tree
species (17 species whose basal areas were, in part, explained by high Ca and soil pH).
All environmental and soil variables explained 37.8% of the variation in bird species
abundance data, with environmental variables explaining 36.0%, soils explaining 0.5%,
and 1.3% being explained by both. Most (67%) of the variation in the bird abundance
data explained by soils was also explained by the vegetation and invertebrate variables,
thereby supporting the hypothesis that the environmental variables were responsible for
the soil-bird relationships.
5.3. Characteristics, Distribution, and Extent of Sensitive
Ecosystems
In the 2008 ISA, it was known that geology (particularly surficial geology) is the
principal factor governing the sensitivity of terrestrial and aquatic ecosystems to
June 2018
5-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
acidification from S and N deposition. Geologic formations with low base cation supply
(e.g., sandstone, quartzite), due mainly to low weathering, generally underlie the
watersheds of acid-sensitive lakes and streams. Bedrock geology has been considered in
numerous acidification studies (Sullivan et al.. 2007b; Vertucci and Eilers. 1993; Stauffer
andWittchen. 1991; Stauffer. 1990; Bricker and Rice. 1989). Other factors contribute to
the sensitivity of soils and surface waters to acidifying deposition, including topography,
soil chemistry, land use, and hydrologic flowpath.
Forests of the Adirondack Mountains of New York, Green Mountains of Vermont, White
Mountains of New Hampshire, the Allegheny Plateau of Pennsylvania, and mountain top
and ridge forest ecosystems in the southern Appalachians are the region's most sensitive
to terrestrial acidification from atmospheric deposition (Section 3.2.4.2 of 2008 ISA).
Recent decreases in acid deposition had been linked to improvements in surface water
chemistry (Appendix 7). However, there remains widespread measurements of ongoing
depletion of exchangeable base cations in forest soils in the northeastern U.S.
(Appendix 4.3.4).
At the time of the 2008 ISA, there had been no systematic national survey to determine
the extent and distribution of terrestrial ecosystem sensitivity to the effects of acidifying
deposition. However, one preliminary national evaluation by McNultv et al. (2007) used
a simple mass balance model and available national databases to estimate forest soil
critical acidifying deposition loads (for wet and dry deposition of S and N) and
exceedances. They found that approximately 15% of forest soils in the U.S. receive
acidifying deposition that exceeds the estimated critical load of wet and dry deposition of
S and N by more than 250 eq/ha/yr (McNultv et al.. 2007). The areas where exceedances
reach this level could be considered to represent those areas that are likely most sensitive
to continued high levels of acidifying deposition.
Since the 2008 ISA, a series of studies have evaluated the characteristics, distribution,
and/or sensitivity of ecosystems to acidifying N and S deposition. For descriptions of
studies that characterized ecosystem sensitivity through critical load and critical load
exceedance determinations, see Appendix 5.5.
5.4. Application of Terrestrial Acidification Models
The models that were used in recent studies to evaluate terrestrial acidification and its
components included the simple mass balance [SMB; (Phelan et al.. 2014; Duarte et al..
2013; Jung et al.. 2013; Whitfield and Watmough. 2012; Forsius et al.. 2010; McNultv
and Boggs. 2010; Nasr et al.. 2010)1. soil texture approximation [STA; (Whitfield et al..
2010bVI. MAGIC (Sullivan et al.. 201 la; Whitfield et al.. 2010a; Whitfield et al.. 2009).
June 2018
5-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
ForSAFE-VEG (McDonnell et al.. 2014a; Sverdrup et al.. 2012). and empirical models
outlined by Spranger et al. (2004) (Krzvzanow ski. 2011). See Appendix 4o for a review
of models.
Phelan et al. (2014) applied the PROFILE model (see Appendix 4.5.1.1) to estimate BCw
to support SMB critical load estimates for 51 hardwood forest sites in Pennsylvania. The
rates of BCw ranged from 119 to 9,245 eq/ha/yr, and were consistent with soil properties
and regional geology. Critical loads ranged from 4 to 10,503 eq/ha/yr. The PROFILE
model estimates by Phelan et al. (2014)were three times larger than those reported for the
same sites by McNultv et al. (2007) who used the clay correlation-substrate method and
SMB models to estimate BCw rates and critical loads, respectively. These PROFILE
model results suggest that the hardwood sites in Pennsylvania may not be as sensitive to
acidifying deposition as previously estimated by McNultv et al. (2007). It should be
noted that BCw rates were not measured for these areas and may be a result of the
different (empirical vs. mechanistic) approaches of each model. As the Phelan et al.
(2014) study only tested and compared the PROFILE model in Pennsylvania, the
researchers recommended applying PROFILE in different regions and ecosystems in the
U.S. to gain a better understanding of the model performance and the degree to which
BCw rates estimated by PROFILE differ from those estimated using the clay
correlation-substrate model.
5.5. Levels of Deposition at Which Effects Are Manifested
Since the 2008 ISA, several studies have evaluated the relationships between N and S
deposition and the growth and physiology of terrestrial organisms and ecosystem
function. In addition, numerous studies have used estimates of critical load exceedances
caused by historic, current, and future N and S deposition levels to characterize the
impacts of acidifying deposition. In this section, these relationships are described as
(1) impacts of elevated N and S deposition, (2) impacts of ambient levels of N and S
deposition, and (3) critical loads and exceedances.
5.5.1. Impacts of Elevated Nitrogen and Sulfur Deposition
A number of studies published since the 2008 ISA have evaluated the impacts of elevated
N and S deposition on trees, herbaceous species, and soil biota (Table 5-3). "Elevated" N
and S deposition in this section refers to additions of N and S above ambient atmospheric
deposition during the time of the specified study. As detailed in Appendix 6. N deposition
may lead to enrichment effects as well as acidification.
June 2018
5-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
In Bear Brook Watershed, ME, Bethers et al. (2009) investigated the effects of
chronically elevated N and S deposition (i.e., 25.2 kg N/ha/yr and 28.8 kg S/ha/yr
additions since 1989) on growth, foliar nutrients, and photosynthetic capacity of sugar
maple saplings. Sugar maple saplings in the treated watershed had higher foliar Al
(+56%), N (+15%), P (+10%), and K (+15%) and lower foliar Ca (-25%) compared to
the nontreated watershed, presumably through influences on soil chemistry. The treated
saplings also had lower photosynthetic capacity, higher N:P ratios, negative correlations
between leaf N and electron transport capacity, and reduced carboxylation capacity,
which suggest nutrient imbalances induced by the elevated N and S deposition. However,
sapling growth was unaffected by the treatments. In another study conducted in Quebec,
Canada, Moore and Houle (2013) observed similar results in their evaluation of the
effects of 8 years of NH4NO3 additions (26 and 85 kg N/ha/yr applied from 2001-2008)
on sugar maple physiology and soil chemistry. Foliar Ca in the high N treatment
decreased by 79% compared to the control and reached 0.24%, the lowest foliar Ca
concentration ever reported for sugar maple. The treatments did not significantly alter
dieback rate or basal area growth. These results corresponded to changes in soil
chemistry; the treatments significantly decreased exchangeable Ca, Mg, Mn, and K in at
least one of the top organic soil layers. The largest differences were observed for
exchangeable Ca between the control and the high N treatment, with the L and the H
layers of the soil organic horizon showing exchangeable Ca decreases of 29 and 72%,
respectively. These results suggest that increased N deposition can strongly affect Ca
nutrition of sugar maple at sites with low base cation saturation.
In West Virginia, Jensen et al. (2014) examined long-term impacts of relatively high N
and S additions (22 years of 169 kg/ha/yr (NH^SO/O on black cherry and yellow poplar
(Liriodendron tulipifera) bole wood Ca, Mg, and Mn concentrations; tree growth; and
basal area. Bole wood Ca concentrations were lower and Mn concentrations higher in
both species on the treated versus nontreated watershed. Growth responses, measured
through relative growth rates of cored trees and changes in basal area, were not
conclusively affected by the treatment.
Jung and Chang (2012) evaluated the impacts of 4 years of elevated N and S deposition
(30 kg N/ha/yr, 30 kg S/ha/yr, and 30 kg N/yr + 30 kg S/ha/yr additions from 2006 to
2009) on a forest composed of trembling aspen (Populus tremuloides), white spruce
{Picea glauca), balsam fir, black spruce (Picea mariana), and paper birch (Betula
papyrifera) in the Athabasca oil sands region of Alberta, Canada. None of the treatments
influenced the growth of the understory plants or soil microbial biomass; understory and
microbial community composition were not measured. However, N increased tree growth
in the N and N + S treatments, indicating N limitation. Nitrogen and S additions
decreased soil exchangeable Ca2+ and Mg2+, and these decreases were attributed to a
5-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
combination of increased tree uptake to support greater growth and increased leaching
with SO42 .
Guv et al. (2013) conducted a greenhouse experiment to assess the potential effects of
acidic deposition on the root system morphology of three endemic species that grow on
sand dunes in the Athabasca oil sands region of Alberta, Canada: Armeria maritima,
Deschampsia mackenzieana, and Stellaria arenicola. Plants were exposed to three pH
treatments (pH 5.6, 5.0, and 4.2) together with additions of 2.61 to 4.67 mg/L of SO42
and 1.95 to 3.51 mg/L NO3 for 55-60 days. There were no statistically significant
differences in plant survival, root length, root surface area, or root tip numbers between
acid treatments. Therefore, current rates of acidifying N and S deposition are not likely a
threat to these species.
Since the 2008 ISA, three studies were identified on the impacts of N and/or S deposition
on soil biota. Payne et al. (2010) examined the impacts of elevated SO42 deposition on
the microbial community in a Scottish peatland and showed that additions of 95 kg
SO42 /ha/yr applied over 18 months decreased the concentrations and percentages of live
amoebae. Abundances of Trinema lineare, Corythion dubium, and Euglypha rotunda
were also significantly reduced. These results suggest the potential importance of SO42
deposition in influencing testate amoebae communities in the peatland soils. In another
study, Chen et al. (2013) evaluated the impacts of S additions on soil microbial and
nematode community composition in an Inner Mongolian grassland in China. Seven
treatments of 28 to 166 mol H+/ha were applied in three doses in the form of sulfuric acid
(H2SO4) from 2009 to 2010. Fungal fatty acids were increased by 49% and fungi:bacteria
ratio increased by up to 120% by the acid additions. Total and bacterial fatty acids were
decreased by the S treatments by up to 47 and 40%, respectively. Acid additions also
impacted the soil nematode community by initially increasing the total number of soil
nematodes, then altering community composition; bacterivorous and fungivorous
nematodes increased, while plant-feeding and omnivorous and carnivorous nematodes
decreased. Hu et al. (2013) examined the impacts of N and S additions on soil microbial
biomass and function in a boreal mixed-wood forest in the Athabasca oil sands region of
Alberta, Canada. Five years of N and S additions (30 kg N/ha/yr and/or 30 kg S/ha/yr)
did not influence soil microbial biomass C and N. However, activities of some
extracellular enzymes in the soil were decreased by the treatments, with greater
(3-glucosidase activity in the N + S treatment than in the S treatment and decreased soil
arylsulfatase activity in the S addition treatment. Thus, the additions of N and/or S
strongly affected soil microbial community functions and enzymatic activities without
changing soil microbial biomass in this boreal forest.
5-29
DRAFT: Do Not Cite or Quote

-------
Table 5-3 Impacts of acidifying nitrogen and sulfur deposition.
Reference
Ecosystem Type/Region
Species
Nitrogen and Sulfur
Deposition/Additions
Description
Bardhan et al. (2012)
Forest
Great Smoky Mountain
National Park, TN, NC
Soil microbes (bacteria)
6 to 41 kg S/ha/yr (modeled)
Soil analyses indicated only minor
differences in bacterial diversity among
sites; the bacterial community did not
change along the gradients of S
deposition.
Bethers et al. (2009)
Forest
Bear Brook watershed,
ME
Sugar maple
Ambient: 8.4 kg N/ha/yr and 14.4 kg
S/ha/yr; elevated: 25.2 kg N/ha/yr
and 28.8 kg S/ha/yr
Treated watershed had a 56% increase in
foliar Al, a 25% reduction in foliar Ca, N
(+15%), P (+10%), and K (+15%). The
treated saplings had lower photosynthetic
capacity, high N:P ratios, and negative
correlations between leaf N and electron
transport capacity, which may indicate
nutrient imbalance.
Boot et al. (2016)
Loch Vale watershed in
Rocky Mountain National
Park, CO
Soil microbes
17 yr of 25 kg N/ha/yr (as NH4NO3)
addition
Long-term fertilization resulted in increased
soil acidity and reduced soil C. Soil
microbial biomass in the fertilized soils was
also lower (22%), the microbial community
was altered through reductions in vesicular
arbuscular mycorrhizae and saprotrophic
fungi, and activity of N degrading microbial
enzymes was decreased.
June 2018
5-30
DRAFT: Do Not Cite or Quote

-------
Table 5-3 (Continued): Impacts of acidifying nitrogen and sulfur deposition.
Nitrogen and Sulfur
Reference	Ecosystem Type/Region	Species	Deposition/Additions	Description
Chen et al. (2013)	Grassland	Soil microbes, bacteria, Sulfuric acid (0, 2.76, 5.52, 8.28, Fungal fatty acids were increased by 49%
Mongolia China	fungi, and nematodes 11.04, 13.8, and 16.56 mol H+/m2) and fungi:bacteria ratio increased by up to
120% by the acid additions relative to the
controls. Total and bacterial fatty acids
were decreased by the S treatments by up
to 47 and 40%, respectively. In the
treatments, bacterivorous and fungivorous
nematodes increased, while herbivorous
and omnivorous + carnivorous nematodes
decreased.
Cleavitt et al. (2011a)
Forest
Acadia National Park, ME
Epiphytic lichens
12 to 18 kg S/ha/yr
Throughfall chemistry influenced bark pH
and the suitability of tree boles as habitat
for lichen; epiphytic lichen species richness
and presence of pollution-sensitive
epiphytes were greater on red maple trees,
which have a higher pH in the bark relative
to red spruce trees.
Dietze and Moorcroft
(2011)
Forest
Eastern and central
portions of the U.S.
267 tree species
organized into 10 PFTs
6 to 16 kg/ha/yr as NO3"; 4 to
30 kg/ha/yr as SO42" (NADP wet)
Mean sensitivity by covariate and PFTs
showed that overall tree mortality was
most sensitive to atmospheric pollutants,
with acid deposition (SO42") showing the
highest sensitivity and N deposition the
third highest sensitivity. Stand DBH
showed the second highest sensitivity.
June 2018
5-31
DRAFT: Do Not Cite or Quote

-------
Table 5-3 (Continued): Impacts of acidifying nitrogen and sulfur deposition.
Reference
Ecosystem Type/Region
Species
Nitrogen and Sulfur
Deposition/Additions
Description
Duarte et al. (2013)
Forest
Northeastern U.S.
21 tree species
256 to 920 eq/ha/yr of N; 242 to
1,154 eq/ha/yr of S (modeled)
Statistically significant negative
correlations between critical load
exceedance and growth (17 species), and
crown density (4 species) were
determined. Positive correlations between
critical load exceedance and declining
vigor (4 species), crown dieback
(4 species), and crown transparency
(7 species) were determined. Species that
were most negatively affected by N and S
deposition included balsam fir, red spruce,
quaking aspen, and paper birch.
Jensen et al. (2014)
Forest
Fernow Experimental
Forest, WV
Yellow poplar and black
cherry
35.5 kg N/ha/yr; 40.5 kg S/ha/yr as
(NH4)2S04 (addition)
Bole wood Ca concentrations were lower
and Mn concentrations were higher in both
species on the treated watershed. Growth
responses were not conclusive and
appeared to differ by species.
Jung and Chang (2012) Forest	Trembling aspen, white 30 kg N/ha/yr, 30 kg S/ha/yr, and Nitrogen addition increased tree growth in
Alberta Canada	spruce, balsam fir, 30 kg N/yr + 30 kg S/ha/yr (additions) the N and N + S treatments. None of the
black spruce, and	treatments affected understory growth or
paper birch	soil microbial biomass. Annual leaching
losses of SO42" were increased by S and
S + N additions. Leaching of base cations
showed a similar trend to SO42" leaching.
Miller and Watmough Forest	Sugar maple	9 to 12.8 kg N/ha/yr; 7.6 to 14.8 kg Sugar maple foliar S and N contents were
(2009)	Southern Ontario	S/ha/yr (ambient)	positively correlated with modeled N and S
Canada	'	deposition. Foliose lichen species richness
was negatively correlated with modeled air
pollution levels (S deposition, N deposition,
and atmospheric ozone).
June 2018
5-32
DRAFT: Do Not Cite or Quote

-------
Table 5-3 (Continued): Impacts of acidifying nitrogen and sulfur deposition.
Reference
Ecosystem Type/Region
Species
Nitrogen and Sulfur
Deposition/Additions
Description
Pannek et al. (2015)
Acidic grassland
Atlantic biogeographic
region of Europe
44 grassland species
within the Violion
caninae alliance
2.4 to 43.5 kg N/ha/yr (ambient)
Out of 44 species studied, 16 were
affected by N deposition, with 12 of them
exhibiting a negative response. Increasing
soil pH and P influenced 24 and
14 species, respectively, predominantly
positively. Species that were negatively
affected by high N deposition and/or high
soil P also showed a negative temporal
trend characterized by short stature and
slow growth.
Payne et al. (2010)
Peatland
Scotland
Soil microbes (testate
amoebae)
Ambient: 5 kg SO42 /ha/yr; elevated:
95 kg S042"/ha/yr
Analysis showed that the SO42" treatment
reduced the concentrations and
percentages of live amoebae, suggesting a
less active community as a result of the
treatment. In addition, abundances of
Trinema lineare, Corythion dubium, and
Euglypha rotunda were significantly
reduced.
Quesnel and Cote
(2009)
Forest
Boreal region of Canada
White spruce
NA
Meta-analysis of foliar results (nutrient
concentrations) from 23 white spruce
forests (natural and nonfertilized plantation
forests) revealed that N deficiencies
occurred in less than 10% of the sites.
Base cation deficiencies identified were
attributed to removals with harvest, higher
requirements of white spruces for K and
Ca, and nutrient deficiencies/imbalances
induced by N additions/deposition.
June 2018
5-33
DRAFT: Do Not Cite or Quote

-------
Table 5-3 (Continued): Impacts of acidifying nitrogen and sulfur deposition.
Reference
Ecosystem Type/Region
Species
Nitrogen and Sulfur
Deposition/Additions
Description
Rose etal. (2016)
Open uplands, open Vascular plants
lowlands, and woodlands
U.K.
NA
Changes in vegetation from 1993-2012 at
a network of plots (504) within
12 Environmental Change Network (ECN)
sites. Significant increases in species
richness were detected at the network
level. Increases in acid-sensitive species
and comparatively little change in
acid-tolerant species were noted. Changes
are consistent with increases in pH
observed and attributed to the large
reductions in acid deposition. Increases in
species diversity were also attributed to
wetter summers and a reduction in soil N
availability at some of the upland locations.
Soule (2011)
Forest
Grandfather Mountain,
NC
Red spruce
NA
Radial growth rates of red spruce
increased through time, and growth rates
were significantly correlated to temperature
(positively), days with precipitation
(negatively), atmospheric CO2 (positively),
and emissions of SOx and NOx
(negatively).
Sullivan etal. (2013)
Forest/Adirondack
Mountains, NY
Sugar maple
750 to 1,120 eq/ha/yr as N + S
(NADP wet, CASTNET dry)
Found that plots without sugar maple
seedlings had higher rates of atmospheric
N + S deposition.
Thomas et al. (2013) Forest	Eastern redcedar	NA
Appalachian Mountains,
WV
Dendroisotopic techniques showed the
recovery of eastern redcedar trees from
decades of S pollution. Analysis provided
evidence for a distinct physiological
response to changes in atmospheric SO2
emissions since 1980.
June 2018
5-34
DRAFT: Do Not Cite or Quote

-------
Table 5-3 (Continued): Impacts of acidifying nitrogen and sulfur deposition.
Nitrogen and Sulfur
Reference	Ecosystem Type/Region	Species	Deposition/Additions	Description
Annual mean and cumulative N deposition
was strongly correlated with decreases in
lichen species richness, decreases in N
sensitive species, and poorer thallus
condition. Cumulative dry deposition ofS
had the best fit to decreases in thallus
condition, poorer community-based S
index values, and the absence of many S
sensitive species.
Al = aluminum; C = carbon; Ca = calcium; CASTNET = Clean Air Status and Trends Network; C02 = carbon dioxide; DBH = diameter at breast height; eq = equivalents;
H+ = hydrogen ion; ha = hectare; K = potassium; kg = kilogram; L = liter; m = meter; mg = milligram; MN = manganese; mol = mole; N = nitrogen; NA = not applicable;
Na2S04 = sodium sulfate; NADP = National Acid Deposition Program; NH4N03 = ammonium nitrate; (NH4)2S04 = ammonium sulfate; N03" = nitrate; NOx = the sum of nitric oxide and
nitrogen dioxide; P = phosphorus; PFT = plant functional types; S = sulfur; S02 = sulfur dioxide; S042" = sulfate; SOx = sulfur oxides; yr = year.
Cleavitt et al. (2015) Forest	Lichens	3 to 8 kg N/ha/yr and 4.5 to 5.2 kg
VT, NH, ME	S/ha/yr
June 2018
5-35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
5.5.2.
Impacts of Ambient Deposition
Five studies that were published since the 2008 ISA have been identified that
documented the impacts of ambient N and S deposition on terrestrial biological
endpoints, including tree species, tree species grouped into plant functional types (PFTs),
lichens, and soil biota (Table 5-3). "Ambient" N and S deposition in this section refers to
the levels of atmospheric N and S deposition experienced during the time of the specified
study.
Dietze and Moorcroft (2011) conducted a large-scale analysis of central and eastern U.S.
that investigated 13 covariates (including climate, air pollutants [N deposition, acid
deposition, and ozone averaged 1994-2005], topography, and stand characteristics) as
predictors of individual tree mortality. Tree species (267 species) were grouped into
10 PFTs based on hardwoods/softwoods, latitude, successional phase (early, mid, and late
successional), and hydrological soil status (i.e., hydric). The researchers found that tree
mortality was most sensitive to stand characteristics and air pollutants. Nine PFTs had
decreased mortality with increased N deposition, and only the northern midsuccessional
hardwoods showing the opposite pattern (i.e., increased mortality with increased N
deposition). Seven PFTs showed large increases in mortality with acid deposition (as
SOr ). with only late successional conifers showing the inverse (i.e., a weak decline in
mortality with increased acid deposition). Mean sensitivity by covariate and PFT showed
that overall tree mortality was most sensitive to atmospheric pollutants, with acid
deposition (SOr ) showing the highest sensitivity and N deposition the third highest
sensitivity. Individual tree DBH showed the second highest sensitivity.
Duarte et al. (2013) evaluated the responses of tree species in 4,057 forest plots in the
northeastern U.S. to the exceedance of critical loads of N and S deposition. Modeled S
and N deposition ranged from 242 to 1,154 eq/ha/yr and 256 to 920 eq/ha/yr,
respectively. Critical loads were exceeded in 45% (calculated using midpoint weathering
rates) of the plots. See Appendix 5.5.3 for a greater description of critical loads and
exceedances from this study. Results from Spearman's rank correlation analyses showed
that the growth of 17 species and crown density of 4 species were negatively correlated
with critical load exceedance (Table 5-4). Positive correlations between critical load
exceedance and declining vigor (three species), crown dieback (four species), and crown
transparency (seven species) were also found (Table 5-5). Crown dieback was considered
the most reliable indicator of forest health. Based on this metric, balsam fir, red spruce,
quaking aspen, red maple, and paper birch were identified as the species most negatively
impacted by N and S deposition.
June 2018
5-36
DRAFT: Do Not Cite or Quote

-------
Table 5-4 Results of Spearman's rank correlation analysis comparing growth
versus critical load exceedance by species for the forested plots in
the northeastern U.S. [from Duarte et al. (2013)1. Modeled sulfur and
nitrogen deposition on the plots ranged from 242 to 1,154 eq/ha/yr
and 256 to 920 eq/ha/yr, respectively. Correlations shown here are
significant at a = 0.05.
Species
Correlation Coefficient
Sample Size (n)
Black spruce (Picea mariana)
-0.44
41
Balsam fir (Abies balsamea)
-0.33
499
Chestnut oak (Quercus prinus)
-0.28
238
Paper birch (Betula papyrifera)
-0.27
574
Scarlet oak (Quercus coccinea)
-0.26
170
Bigtooth aspen (Populus grandidentata)
-0.25
280
Eastern white pine (Pinus alba)
-0.19
1,464
Black oak (Quercus velutina)
-0.18
257
White ash (Fraxinus americana)
-0.18
1,256
Sweet birch (Betula lenta)
-0.12
460
Yellow birch (Betula alleghaniensis)
-0.12
669
Northern red oak (Quercus rubra)
-0.11
1,533
Red maple (Acer rubrum)
-0.11
3,861
Red spruce (Picea rubens)
-0.09
616
American beech (Fagus grandifolia)
-0.08
1,449
Black cherry (Prunus serotina)
-0.07
1,026
Eastern hemlock (Tsuga canadensis)
-0.07
1,055
Pignut hickory (Carya glabra)
0.44
31
Norway spruce (Picea abies)
0.38
148
Yellow poplar (Liriodendron tulipifera)
0.35
47
White spruce (Picea glauca)
0.24
97
eq = equivalents; ha = hectare; yr = year.
Source: Duarte et al. (2013).
June 2018
5-37
DRAFT: Do Not Cite or Quote

-------
Table 5-5 Results of Spearman's rank correlation analyses comparing tree
vigor, crown density and dieback, and canopy transparency versus
critical load exceedance by species for the forested plots in the
northeastern U.S. [from Duarte et al. (2013)1. Modeled sulfur and
nitrogen deposition on the plots ranged from 242 to 1,154 eq/ha/yr
and 256 to 920 eq/ha/yr, respectively. Correlations shown here are
significant at a = 0.05.

Vigor

Crown Density
Crown Dieback
Canopy Transparency
Species
Correlation
Coefficient
Sample
Size (n)
Correlation Sample
Coefficient Size (n)
Correlation
Coefficient
Sample
Size (n)
Correlation
Coefficient
Sample
Size (n)
American
beech
0.13
433

0.12
502
0.30
502
Ash
0.22
189





Balsam
fir


-0.23 161
0.56
161
0.18
161
Black
cherry
0.33
47





Eastern
hemlock


0.36 82
-0.24
148


Fir





-0.36
33
Northern
red oak





0.27
208
Paper
birch


-0.29 125
0.13
300


Quaking
aspen


-0.56 61
0.44
61
0.40
61
Red
maple


-0.16 337
0.10
767
0.19
767
Red
spruce


0.20 115
0.43
168
0.61
168
Sugar
maple
0.06
3,408

-0.04
3,466
0.12
3,467
White
ash
-0.32
49





eq = equivalents; ha = hectare; yr = year.
Source: Duarte et al. (2013)
June 2018
5-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Sullivan et al. (2013) evaluated relationships between deposition and sugar maple
regeneration in the Adirondack Mountains. Total N and S deposition ranged between 750
and 1,120 eq/ha/yr, and plots that received higher rates of N (median of approximately
560 eq/ha/yr), S (median of approximately 480 eq/ha/yr), and N + S (median of
approximately 1,040 eq/ha/yr) deposition had no sugar maple seedlings.
Soule (2011) sampled red spruce trees within a high elevation red spruce-Fraser fir (Abies
fraseri) forest on Grandfather Mountain, NC between 2006 and 2008, and used
dendrochronology to evaluate the influences of climate (precipitation, temperature, and
CO2 concentrations) and emissions of SOx and NOx on growth. Radial growth rates of
red spruce increased through time and were positively correlated to temperature and
atmospheric CO2, but were negatively correlated to emissions of SOx and NOx and
number of days with precipitation in a year.
Miller and Watmough (2009) evaluated hardwood plots along air pollution (N, S, and
ozone), soil acidity, and climate gradients in Ontario, Canada, and found that foliose
lichen species richness was negatively correlated with modeled air pollution levels (S
deposition, N deposition, and atmospheric ozone). In the same study, however, no
relationship was seen in canopy condition and ground vegetation richness and diversity.
Bardhan et al. (2012) evaluated bacterial diversity along a soil and S deposition gradient
in high elevation spruce-fir forests in GSMNP. Modeled S deposition ranged from 6 to
41 kg S/ha/yr. However, neither bacterial diversity nor community composition changed
along the gradient of S deposition. The bacterial community on all sites was dominated
by members of the phyla Actinobacteria, Acidobacteria, Planctomycetes, Proteobacteria,
and Chloroflexi. Species from these phyla are often found in highly acidic environments
such as acid-mine drainage and sphagnum bogs. Therefore, these analyses suggested that
despite reductions in acid deposition, the soil conditions in these GSMNP sites were still
acidic and had not yet reached a threshold suitable for nonacidophilic bacterial
communities.
Pannek et al. (2015) used a data set from seminatural grasslands found on acidic soils
along an N deposition gradient (2.4 to 43.5 kg N/ha/yr) in northwestern Europe to
examine the response of species frequency to N and other factors including soil
(0-10 cm) P, pH, NH4+, and NO3 and geographical, climatic, and management factors.
Out of 44 studied species, 16 were affected by N deposition, with 12 of them exhibiting a
negative response (see also Appendix 6.2.5 and Appendix 6.3.5).
5-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
5.5.3.
Critical Loads and Exceedances
Since the 2008 ISA, critical load evaluations have been conducted in the northeastern
forests of the U.S., hardwood forests in Pennsylvania, forests in western Canada, and
arctic, subarctic, alpine, subalpine, and boreal systems. Studies also simulated target
loads and critical loads and exceedances into the future under different deposition and
climate scenarios, and evaluated the sensitivity of critical load estimates to varying
environmental conditions and forest management and uncertainties in critical load
estimates. See Appendix 4.6.2.1 for more information on critical loads based on soil
acidification and Appendix fo for information on N critical loads in terrestrial systems.
Duarte et al. (2013) calculated critical loads of N and S for 4,057 forested plots in the
northeastern U.S. using the steady-state SMB model (see also Appendix 4.5.1.2')and soil
solution Be:A1 of 10.0 as the chemical indicator and threshold. In addition, the study
evaluated the influence of three different BCw rates on critical loads: minimum,
midpoint, and maximum. These BCw rates were determined using the clay
correlation-substrate method and the range of U.S. Department of Agriculture (USDA)
Natural Resources Conservation Service (NRCS) Soil Survey Geographic Database
(SSURGO) values for each soil series. Critical loads using the midpoint weathering rates
were found to range from 11 to 6,540 eq/ha/yr (over 80% within the range of
850-2,050 eq/ha/yr), and in comparisons with deposition estimated using the ClimCalc
model, were exceeded in 98% (calculated using minimum weathering rates), 45%
(calculated using midpoint weathering rates), and 15% (calculated using maximum
weathering rates) of plots. Similarly, Phelan et al. (2014) calculated critical loads of N
and S deposition for 51 hardwood forests in Pennsylvania using the SMB model, soil
solution Bc:Al of 10.0 as the chemical indicator, and the PROFILE model to estimate
BCw rates. They found that critical loads ranged from 4 to 10,503 eq/ha/yr, and that
critical loads at 53% of the sites were exceeded by the 2002 N and S deposition.
Krzvzanowski (2011) modeled deposition and soil acidification critical loads in
northeastern British Columbia, Canada using empirical methods described in Spranger et
al. (2004). Slow weathering of shale- and mudstone-derived feldspars, micas, and quartz
placed the soils of the study area in the "high-sensitivity" class of critical loads. Soil
acidification critical load was estimated to be 200 eq/ha/yr. Neither S nor N deposition
exceeded this critical load. The combined S and N deposition at the dominant point of
reception (i.e., near Blueberry River First Nation) was estimated to be 73.5 eq/ha/yr.
Estimates of S and N critical loads and exceedances for upland soils were generated for
the Georgia Basin in British Columbia, Canada using the SMB model and zero base
cation depletion as the chemical criterion and threshold (Nasr et al.. 2010). The objective
of the study was to evaluate critical load and exceedance calculations in the context of
June 2018
5-40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
sustainable forest and soil management policies by using "no further changes in soil base
saturation" as the critical load criterion. Critical loads were found to range between 140
and 4,000 eq/ha/yr, and were generally lowest on ridge tops and increased towards the
valleys. Critical load exceedance ranged from 13% of the basin when exposed to wet
deposition to 32% under modeled total (wet and dry) deposition. With continued N and S
deposition, significant portions of the basin were predicted to experience
exceedance-enhanced base cation depletion rates greater than 200 eq/ha/yr.
Forsius et al. (2010) determined critical loads and exceedances for terrestrial ecosystems
in the arctic and subarctic regions (with latitudes north of 60°) for 1990, 2000, 2010, and
2020 using the SMB model and two different chemical indicators (soil solution Al:Bc of
1.0 and ANCle of 0.0). The Al:Bc indicator was hypothesized to protect against fine root
damage, and the ANCle indicator was hypothesized to preserve existing soil base cation
pools. The 2020 deposition scenario was based on the maximum feasible reduction
(MFR) emissions for 2020. Critical load estimates were generally comparable among the
three main regions (northern Europe, Russia, and North America) and were influenced by
the chemical indicator and associated threshold. The Al:Bc of 1.0 was less stringent than
ANCle of 0.0, with the median critical loads for the two indicators being 700 and
300 eq/ha/yr, respectively. In North America, the lowest critical loads occurred in eastern
Canada above latitudes of 60°. In general, low critical loads were found in areas with low
weathering rates associated with coarse soils on acidic parent material. Critical loads in
North America (above latitudes of 60°) were not exceeded by any of the deposition levels
for any of the years: The lowest critical load was 130 eq/ha/yr, while the maximum rates
of N and S deposition in North America were estimated to be 30 to 40 eq/ha/yr.
Simkin et al. (2016) quantified the effect of N deposition on species richness in the
continental U.S. using data from over 15,000 plots. N deposition had a strong effect on
species richness, but this effect differed between closed canopy and open ecosystems
(i.e., forests and nonforest). In nonforested ecosystems (grasslands, deserts, shrublands,
subalpine ecosystems), there was a positive relationship between N deposition and
herbaceous species richness at low rates of N deposition, then a decrease in species
richness with higher rates of N deposition over a threshold of 8.7 kg N/ha/yr. This study
focused N enrichment, but highlighted interactions with soil acidification. Vegetation on
acidic soils was more susceptible to species loss under elevated N. See Appendix 62 and
Appendix 6.5.3 for more information on this study.
5-41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
5.5.3.1. Target and Future Critical Loads/Critical Load Exceedances
Target loads of S deposition were calculated for 44 watersheds and extrapolated to
I,320	acid-sensitive watersheds in the Adirondacks using MAGIC (see also
Appendix 4.5.1.4) and different soil chemical indicator and thresholds of base saturation
(5 and 10%), soil solution Bc:Al, and Ca:Al [1.0 and 10.0; (Sullivan etal.. 201 la)I. In a
comparison of target loads (out to Years 2050 and 2100) and the 2002 deposition, only
II.6	to 13.5% of the watersheds were simulated to be in exceedance of target loads to
protect soil base saturation to 5%. For target loads to protect soils to a base saturation of
10%, 79.7 to 87.5% of the watersheds were in exceedance. For soil solution Bc:Al, 7.8
and 98.1% of watersheds were exceeded by the 2002 deposition for target loads to protect
soil solution ratios of 1.0 and 10.0, respectively. For soil solution Ca:Al, 44.1 to 58.2% of
watersheds experienced exceedances of target loads associated with a soil solution ratio
of 1.0, and 98.1% of watersheds with target loads to protect Ca:Al ratios of 10.0 were
exceeded. Further investigations revealed that 58.2, 85.7, and 93.6% of watersheds could
not obtain threshold values of 10% base saturation, Bc:Al = 10, or Ca:Al = 10,
respectively, even if acidic deposition was held at zero, thereby demonstrating that these
chemical indicator threshold values were not useful for target load calculations using
MAGIC in the Adirondack Mountains.
Sverdrup et al. (2012) and McDonnell et al. (2014a) used the ForSAFE-VEG model (see
also Appendix 4.5.1.3) to evaluate potential long-term effects of climate change and
atmospheric N deposition on alpine/subalpine ecosystems. Critical loads in both studies
were defined as the amount of N deposition to protect against a change in plant
biodiversity of 5 to 10 Mondrian (M) units (i.e., 5-10% change in plant species cover).
Sverdrup et al. (2012) focused on a "generalized" alpine/subalpine site in the Rocky
Mountain National Park, and simulated plant responses to a future climate (IPCC
scenario A2) and four levels of N and S deposition (preindustrial background, Clean Air
Act [CAA] controls, no CAA controls, and no CAA controls + high deposition). Soil
base saturation decrease to less than 1% and soil solution Bc:Al were predicted to be less
than 10.0 after Year 2100, with the futures of no CAA emission controls and no CAA
controls + high deposition (indicating soil acidification). Future plant species coverages
were predicted to change in successively greater amounts in response to the altered
climate, CAA emission controls, no CAA emission controls, and no CAA emission
controls + high deposition. Critical loads of N deposition were calculated (based on a
change of 5 M) to be 1 kg N/ha/yr. Critical loads related to S deposition were not
discussed. All future N deposition scenarios (except preindustrial background N) were
simulated to result in critical load exceedance. McDonnell et al. (2014a') found similar
results in their study that evaluated the response of plants on a site in the Loch Vale
watershed, CO to 100 different scenario combinations of N deposition, precipitation, and
June 2018
5-42
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
temperature. The estimated critical load of N to protect against future (average of
Years 2010-2100) change in biodiversity (10 M) was estimated to be between 1.9 and
3.5 kg N/ha/yr, depending on the temperature increase scenario. Current deposition levels
were found to exceed the critical load. Future increases in temperature were forecasted to
substantially impact plant community composition beyond the predicted changes in
response to N alone; N deposition was forecasted to result in approximately a 10- to
25-M change by 2100, whereas a +4.6°C increase in temperature was forecasted to result
in approximately a 38- to 48-M change by 2100. In both studies, plant community
response appeared to be attributed to N enrichment or eutrophication, as McDonnell et al.
(2014a) stated that the critical load adopted in their study was much lower than critical
loads to protect against NOs leaching and soil acidification determined by other studies.
5.5.3.2. Sensitivity of Critical Load Estimates to Forest Management and
Environmental Stresses
Influences of Forest Management
Removal of N and base cations with tree harvesting can be included in the calculation of
critical loads. Several recent studies evaluated the amounts of base cations removed
through different harvesting practices and determined that base cation removal can be
substantial and contribute to site acidification.
Duchesne and Houle (2008) determined base cation budgets in a managed boreal balsam
fir forest in Quebec, Canada according to six different scenarios, including two
harvesting scenarios (whole-tree and stem-only harvesting), and three scenarios of
mineral weathering. Whole-tree harvesting was found to remove twice as much Ca (1,358
vs. 664 mol/ha) and K (483 vs. 200 mol/ha) as stem only. In contrast to Ca and Mg,
immobilization of K within tree biomass (69 mol/ha/yr) was the main pathway of K
losses from the soil exchangeable reservoir, being five times higher than losses via soil
leaching (14 mol/ha/yr). The amounts of K contained within the aboveground biomass
and the exchangeable soil reservoir were 3.3 and 4.2 kmol/ha, respectively. Whole-tree
harvesting was estimated to remove 44% of the K that was readily available for cycling
in the short term.
Similar results were found by Iwald et al. (2013) in their evaluation of the removal of
base cations with the harvesting of tree stumps and logging residues for biofuels in
Sweden. Their study evaluated three levels of harvesting intensity that varied in the
amounts of stump and logging residues left on site (0-40%). Results from the study
indicated that harvesting of stumps and logging residues constituted 13 to 24% and 27 to
June 2018
5-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
45% of total (stumps + stem wood + logging residues) base cation extraction,
respectively, depending on harvest intensity and tree species. The higher acidifying effect
of logging residue removal was explained by the higher contents of base cations in
needles and branches compared to stem wood. In a comparison between total net cation
extraction by tree harvesting and maximum levels of current acid deposition, the
acidifying effect of pine harvesting was found to be 57 to 108% of that of acid
deposition, the acidifying effect of spruce harvesting was 114 to 263%, and the acidifying
effect of birch harvesting was 60 to 171%.
In the northeastern U.S., Lucas et al. (2014) evaluated base cation extractions under three
different management scenarios that varied in management intensity and conservation
focus. The MaxGrowth management option, which involved site scarification, intensive
fertilization (amonium nitrate), and short rotation lengths (60-100 years), resulted in
roughly 50 to 100% greater base cation removal than that associated with the other three
management scenarios. Additionally, removal of tree tops, branches, and stumps
consistently resulted in removal rates of Ca, Mg, and K three to four times greater than
conventional stem-only harvests.
Johnson et al. (2015a) used mass balance calculations to evaluate the impact of
harvesting on ecosystem balances of Ca, Mg, K, and acidification of forest soils at
40 sites in Ireland. Three harvesting scenarios were evaluated: stem-only harvest (SOH),
stem plus branch harvest (SBH), and whole-tree harvest (WTH). Mass balances for Ca,
Mg, and K were determined based on the difference between long-term inputs
(atmospheric deposition plus mineral weathering) and outputs (biomass removal plus
leaching losses). Soil acidification was calculated using a simplified acidity balance of
inputs (base cation and Na deposition and base cation weathering) and outputs (base
cation uptake, S deposition, and Cl~ deposition). Under the SOH and SBH management,
inputs of Ca, Mg, and K were predicted to be sufficient to meet outputs. Atmospheric
deposition was the most important source of Ca and Mg input. For K, inputs from soil
weathering were as important as deposition. Under the WTH scenario, Ca output
exceeded input at 19 of the 40 study sites. However, the differences were small relative to
the sizes of the Ca pools; at the 19 sites, exchangeable pools could support Ca removal
with WTH for a median of 220 years. Mg and K removal with WTH was balanced by
inputs through deposition and soil weathering. In contrast, for soil acidification budgets,
base cation removal with all scenarios of harvesting was much greater than that generated
by soil weathering, suggesting that soils will become acidified over the long term.
However, there was considerable uncertainty around the calculation of base cation fluxes.
For Ca balances, confidence intervals spanned positive and negative values at so many
sites that it was not possible to predict the balance of Ca budgets. In addition, uncertainty
5-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
in flux calculations was particularly important for K, as soil exchangeable pools were
small and could be depleted within one or two rotations.
Zetterberg et al. (2014) evaluated the long-term effects of whole-tree harvest (WTH) on
soil and stream water acidity in three forested catchments dominated by Norway spruce
in Sweden. Potential influences of varying the amounts of logging residues, Ca
concentrations in tree biomass, and site productivity on the model predictions were also
evaluated through a sensitivity analysis. The MAGIC model was used to simulate
changes in forest soil exchangeable Ca pools and stream water ANC from 1850 to 2100,
with WTH occurring in 2020. Large depletions in soil Ca supply and a reversal of the
positive trend in stream ANC were predicted for all three catchments sites after WTH.
However, the magnitude of impact on stream ANC varied depending on site and the
concentrations of mobile strong acid anions. Varying the tree biomass Ca concentrations
was found to have the largest impacts on modeled soil and stream chemistry. Site
productivity was the second most influential variable, and changing the amount of
harvest residues left on site only marginally affected soil exchangeable Ca and stream
water ANC. The results from this study suggest that future research should concentrate
on minimizing uncertainties in tree biomass Ca concentrations and performing studies on
biological feedback mechanisms that can increase Ca availability in the soils.
Influences of Environmental Stresses
Climate and other environmental stresses may also directly or indirectly alter ecosystem
parameters that are used in the SMB model to determine critical loads. The potential
influences of climate were examined by McNultv and Boggs (2010) in their case-study
evaluation of red spruce stands in western North Carolina that were experiencing low
versus high rates of mortality associated with a pine beetle outbreak. There were positive
relationships between site fertility (forest floor and soil measurements, foliar N
concentrations, and Mg:N ratios) and red spruce mortality. Annual basal area growth of
red spruce on more fertile plots was more sensitive to drought than on less fertile sites.
Based on these observations, critical loads of acidity could change as a result of episodic
stress (e.g., drought, insect infestations, frost injury, cold tolerance, etc.), and increased
growth due to higher fertility and/or nutrient imbalances caused by acidifying deposition
may make red spruce more susceptible to environmental stresses. Critical loads of N and
S, therefore, might be lower when environmental stresses are present.
5-45
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
5.5.3.3.
Uncertainties in Critical Load Estimates
Reinds and de Vries (2010) evaluated the uncertainties in critical and target loads of S
and N for 182 European forest soil plots using the very simple dynamic (VSD) model
(see also Appendix 4.5.2.2V Target loads were defined as the deposition that leads to a
desired chemical state of the ecosystem in a given future year. The VSD model was
calibrated using Bayesian prior probability functions for model parameters based on
literature data, data from 200 Dutch forest sites and from simulated denitrification rates
from a detailed ecosystem model, which improved the model fit to observed soil
measurements. Critical loads determined by the calibrated model varied by chemical
criterion. Minimum critical loads of N ranged between 181 (5th percentile) and
1,606 eq/ha/yr (95th percentile), maximum critical loads of N ranged between 502
(5th percentile) and 31,247 eq/ha/yr (95th percentile), and maximum critical loads of S
ranged between 4 (5th percentile) and 9,670 eq/ha/yr (95th percentile), with the critical
loads based on Al:Bc = 1 being higher than those determined using the ANC = 0 eq/ha/yr
criterion. Uncertainty analysis also showed that the main drivers of uncertainty were
largely dependent on the chemical criterion used in the critical and target load
calculations. Base cation weathering, deposition, and the parameters describing the H-Al
equilibrium in the soil solution were the main sources of uncertainty in the estimates of
maximum critical loads for S (CLmax[S]) based on the Al:Bc criterion of 1.0, and
uncertainty in CLmax(S) based on ANC was completely determined by base cation
inputs. The denitrification fraction was the most important source of uncertainty for the
maximum critical loads of N (CLmax[N]). Calibration of VSD reduced the levels of
uncertainty for all critical loads and criteria. After calibration, the coefficient of variation
(CV) for CLmax(S) was below 0.4 for almost all plots, and target loads were not needed
in any of the simulations for 40% of the plots. According to the noncalibrated model,
there was a positive probability for the need of a target load for almost all plots.
5.6. Climate Modification of Ecosystem Response
The effect of acidifying deposition on terrestrial ecosystems can be modified by climate
shifts in temperature and precipitation. Appendix 13 provides an overview on this topic
and Appendix 4.7.1. provides information on climate modification of soil acidification.
Warmer temperatures increase decomposition and nitrification. Nitrification will also
increase with increased N supply caused by increased weathering or decomposition
(Booth et al.. 2005). The process of nitrification generates protons that increase the rate
of nitrate and base cation leaching to drainage waters (Murdoch et al.. 1998). The
combined increase of NO3 leaching and loss of base cations has the potential to magnify
June 2018
5-46
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
acidification in forest soils (Fernandez et al.. 2003). Soil weathering is typically the key
buffer to acidic deposition (Li and McNultv. 2007). and while weathering is increased by
both soil temperature and soil moisture (Gw iazda and Broecker. 1994). it is unclear
whether any future change in the magnitude of temperature and precipitation will be
enough to alter base cation supply or influence the acid-base balance of sensitive
ecosystems. Furthermore, it is unclear whether increased supply of N in soils from either
deposition, increased decomposition, or increased nitrogen fixation may negate the
ameliorative effect of enhanced weathering. Some studies show that climate change will
mitigate acidification through increased weathering (Bclvazid et al.. 201 la), while others
show that climate change will aggravate acidification although increased nitrification
outpacing enhanced weathering (Wu and Driscoll. 2010). In general, increased
temperature and precipitation will likely enhance inputs of buffering agents from
weathering and deposition, but also increase inputs of acidifying agents from deposition
and enhanced N cycling. The relative sensitivity of these opposing processes to a given
change in climate remains unresolved.
The mountainous regions of the Eastern U.S. are especially interesting to study because
of the both acidifying deposition and climate change interact in these areas. Recently,
Wason etal. (2017) studied the responses of red spruce and balsam fir to acidic
deposition and trends in climate on Whiteface Mountain in New York using tree ring
analysis in forests plots along an elevational gradient. They found that both species
increased growth with increased precipitation pH. Red spruce growth appeared to
increase growth with a warming climate and balsam fir did not. Despite the changes in
growth due to precipitation chemistry and climate, the authors did not detect changes in
the distribution in the spruce-fir forest and perhaps this is a longer-term process. The
study demonstrated the complexity of forest response as multiple environmental factors
change as these forests recover form acidifying deposition.
Koo et al. (2014)used the Annual Radial Model (ARIUM) to investigate projected
climate change effects and changing air pollution on red spruce growth in the Great
Smoky Mountain national park. The model estimated that high elevation (<1,700 m) red
spruce growth would decline 10.8% when a climate change interacted with a 10%
increase in air pollution. However, growth increased by 8.4% when air pollution
decreased 10% in air pollution with climate change. In contrast, low elevation red spruce
growth decreased with future climate change with decreased, increased and no change in
air pollution.
5-47
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
5.7.
Summary
Since publication of the 2008 ISA, the overarching understanding of terrestrial
acidification has not appreciably changed. More recent research has confirmed and
strengthened this understanding and provided more quantitative information, especially
across the regional landscape. A number of studies have evaluated the relationships
between soil chemistry indicators of acidification and ecosystem biological endpoints
(see Table 5-6). Soil chemistry indicators examined in recent literature include
exchangeable base cations, soil pH, exchangeable acidity (H+ and Al), exchangeable
Bc:Al ratio, base saturation, and Al concentrations. Biological endpoints included in the
evaluations consisted of physiological and community responses of trees and other
vegetation, lichens, soil biota, and fauna.
5.7.1. Physiology and Growth
The physiological effects of acidification on terrestrial ecosystems in the U.S. were well
characterized at the time of the 2008 ISA. Consistent and coherent evidence from
multiple species and studies in 2008 showed that the biological effects of acidification on
terrestrial ecosystems were generally attributable to physiological impairment caused by
Al toxicity and decreased ability of plant roots to take up base cations (Section 3.2.2.3 of
the 2008 ISA). Acid deposition can also leach membrane-associated stores of Ca from
young red spruce needles, which affects membrane stability and freezing tolerance. New
information since the 2008 ISA has supported these conclusions (Appendix 5.2).
including further studies on the impact of acidifying deposition on sensitive tree species,
such as sugar maple and red spruce. Much of the new evidence for the negative effects of
acidifying deposition on these species comes from Ca addition experiments, in which the
addition of Ca has alleviated many of the negative plant physiological and growth effects.
Consistent with the findings of the 2008 ISA, the body of evidence is sufficient to infer
a causal relationship between acidifying N and S deposition and the alteration of the
physiology and growth of terrestrial organisms and the productivity of terrestrial
ecosystems.
In the 2008 ISA, acidifying deposition, in combination with other stressors, was found to
be a likely contributor to physiological effects that led to the decline of sugar maple trees
living at higher elevation in some portions of the eastern U.S. that have geologies
dominated by sandstone or other base-poor substrate, and that have base-poor soils.
Studies since the 2008 ISA support these findings (see Appendix 5.2.1.1). For example,
recent field studies demonstrated relationships between soil chemical indicator threshold
values and tree responses. Substantial declines in sugar maple regeneration were found at
June 2018
5-48
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
soil base saturation levels <20%, which is consistent the range reported in the 2008 ISA.
In new studies, sugar maple showed positive growth and regeneration responses to
increasing exchangeable base cations, base saturation and soil pH, and negative
relationships with increasing exchangeable Al. Additional studies indicated that growth,
regeneration and physiological responses of sugar maple to soil conditions created by
acidifying N and S deposition were reversed or ameliorated by Ca additions. The
responses to Ca additions included increased growth and regeneration, decreased crown
dieback, increased foliar chlorophyll, and decreases in foliar metabolic indicators of
physiological stress.
The 2008 ISA reported that changes in soil chemistry, such as depletion of soil base
cations, increasing Al concentration, and leaching of base cations into drainage water,
have contributed to physiological stress, high mortality rates, and decreasing growth
trends of red spruce trees. New information since the 2008 ISA from Ca addition studies
supports the conclusion that depletion of base cations contributed to these effects in trees
(Appendix 5.2.1.2). Foliar biomass and physiological responses of red spruce to soil
conditions created by acidifying N and S deposition were reversed or ameliorated by Ca
additions. The responses included higher foliar antioxidant activity in the winter,
significantly greater foliar cold tolerance, and higher levels of foliar metabolic
compounds that indicate an increased tolerance of environmental stresses. (Schabcrg et
al.. 2011; Halman et al.. 2008).
In the 2008 ISA, there was a limited amount of information on acidification effects on
flowering dogwood. Since the 2008 ISA, no additional information on dogwood has been
published; however, other tree species that have been evaluated in studies relating soil
chemistry to tree physiology include yellow birch, white ash, American beech, black
cherry, northern red oak, hickories, American basswood, and eastern hophornbeam.
These species were reported to vary in their sensitivities to soil conditions associated with
acidifying deposition. Data are insufficient to draw general conclusions for other species.
New information was also published on the recovery of eastern redcedar since the 1980s
as SO2 emissions declined. However, it is unclear whether this recovery was from acid
deposition, direct SO2 effects, or a combination of each (Appendix 5.2.1.3).
New studies since the 2008 ISA have also added new information about acidifying
deposition on forest understory vegetation, grasslands, lichen communities, and higher
trophic-level organisms (snails and salamanders). These studies are not as numerous as
those from the decades of tree research; however, results of these studies support the
conclusions of the 2008 ISA regarding the effects of acid deposition on terrestrial
ecosystems.
5-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
5.7.2.
Biodiversity
Consistent with the findings of the 2008 ISA, the body of evidence is sufficient to infer
a causal relationship between acidifying N and S deposition and the alteration of
species richness, community composition, and biodiversity in terrestrial ecosystems.
The physiological effects of acidifying deposition can result in changes in species
composition and biodiversity whereby sensitive species are replaced by more tolerant
species. For example, sugar maple was found to have greater growth and seedling
colonization with increasing soil cation availability, and American beech was more
prevalent on soils with lower levels of base cations—locations where sugar maple does
less well (Appendix 5.2.1.3.1V As noted above, studies have found increased
regeneration of sugar maple with Ca additions and less regeneration with increasing
exchangeable Al. Soil acid-base chemistry was found to be a predictor of understory
species composition. Fifty understory species were associated with the basic end of a soil
pH gradient, and these species could have value as indicators of sites with high base
cation status and potentially suitable habitat for acid-sensitive species like sugar maple.
In another set of studies, soil acid-base chemistry predicted and was correlated with soil
biota diversity and community composition. Proportions of soil bacteria, fungi, and
nematodes were found to be correlated to soil pH and Al concentrations. Fungi and
nematodes were more abundant in acidic soils. Ca additions resulted in a change in soil
bacterial community composition, and the bacterial community structure was found to be
significantly correlated with soil exchangeable Ca, pH, and P (Appendix 5.2.4).
5.7.3. National-Scale Sensitivity and Critical Loads
Sensitivity of soils to acidifying deposition is discussed in detail in Appendix 4 and
summarized in Chapter 1.5. In general, surficial geology is the principal factor governing
the sensitivity of terrestrial ecosystems soil to acidification from S and N deposition.
Other factors contribute to the sensitivity of soils to acidifying deposition, including
topography, soil chemistry, and land use. Forests of the Adirondack Mountains of New
York, Green Mountains of Vermont, White Mountains of New Hampshire, the Allegheny
Plateau of Pennsylvania, and mountain top and ridge forest ecosystems in the southern
Appalachians are the regions most sensitive to terrestrial acidification from atmospheric
deposition (Section 3.2.4.2 of the 2008 ISA). Sensitive ecosystems can also be
characterized by presence of acid-sensitive soils and plant species (Appendix 5.3).
Models used to determine critical loads of acidifying deposition included: SMB, STA,
MAGIC, ForSAFE-VEG, and empirical models. Several models and extrapolation
methods to estimate BCw rates were also investigated. The PROFILE model was
June 2018
5-50
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
evaluated as a model to estimate soil BCw rates to support estimates of SMB critical
loads in the U.S. In general, recently published models used soil solution Bc:Al of 10.0 as
an indicator to estimate critical loads in North America. These models are described in
more detail in Appendix 4^5.
Sensitivities of ecosystems to ambient N and S deposition were also characterized
through critical loads and exceedances (Appendix 5.5). Calculated critical loads for forest
sites based on the soil solution Bc:Al of 10.0 in the northeastern U.S. ranged from 11 to
6,540 eq/ha/yr (eq quantifies the supply of H+ ions available for acid-base reaction,
allowing the acidifying effects of N and S deposition to be combined into the same unit),
and 15-98% of these critical loads were exceeded by N and S deposition. In this region,
correlation analyses showed that the growth of 17 species and crown density of 4 species
were negatively correlated with critical load exceedance. In Pennsylvania, critical loads
based on the soil solution Bc:Al of 10.0 for hardwood forests ranged from 4 to
10,503 eq/ha/yr and were exceeded by the 2002 N and S deposition in 53% of the plots.
In comparison, critical loads for terrestrial ecosystems in the arctic and subarctic regions
of North America were not exceeded by estimated deposition in the Years 1900, 2000,
2010, and 2020. For these high latitude ecosystems, the lowest critical load was
130 eq/ha/yr, while the maximum N and S deposition was 30 to 40 eq/ha/yr.
In western Canada, critical loads ranged from 40 to 4,000 eq/ha/yr depending on the
study and location. Several studies evaluated the influence of BCw rates, soil chemical
indicators and thresholds, N retention, tree species-specific base cation uptake, and/or
bulk (i.e., wet) versus total deposition on critical load estimates. All of these parameters
were found to influence critical load and exceedance determinations, thereby
demonstrating the uncertainties and sensitivities associated with critical load estimates.
5-51
DRAFT: Do Not Cite or Quote

-------
Table 5-6 Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Bardhan et al.
(2012)
Forest
Great Smoky
Mountain
National Park,
TN, NC
Soil
microbes
(bacteria)
6 to 41 kg
Soil pH and
S/ha/yr
exchangeable
modeled
Ca and Al, CEC
deposition
and base
classes
saturation (O-,
(Weathers et
A-, and
al.. 2006)
B-horizons)
14.57 to
Exchangeable
19.7 kg/ha/yr as
Ca and pH (Oa-
wet NO3"; 17.44
and B-horizons)
to

29.09 kg/ha/yr

as wet SO42",

modeled (Ito et

al.. 2002)

1990-1999

Only minor differences in bacterial diversity among sites; the bacterial
community did not change along the gradients of S deposition, soil pH,
or exchangeable Ca:AI ratio. High elevation sites remain acidic and
have not yet reached a threshold suitable for nonacidophilic bacterial
communities.
Beier et al. (2012)
Forest
Adirondack
Mountains, NY
(12 upland
hardwood
forests)
NA
Increasing trends in snail community richness and abundance, live
biomass of red-backed salamanders, and canopy tree basal area with
increasing soil Ca. Land snail species richness and abundance were
positively correlated with Oa-horizon Ca and negatively correlated to
SO42" deposition. Salamander communities changed continuously
along the Ca gradient. Several known calciphilic species of snails and
plants were found only at the highest Ca sites. The proportion of basal
area attributed to standing dead trees decreased significantly with
Oa-horizon Ca.
Bilodeau-Gauthier
et al. (2011)
Forest
Quebec lakes
network,
Quebec,
Canada
(6 watersheds)
Sugar
maple
NA
Soil pH, base	Tree growth was positively correlated to concentrations of base cations
saturation,	(Ca, K, and Mg) in wood and soil, and negatively correlated to
exchangeable	concentrations of acidic metals in wood (Al, Mn, and Cd) and soil (H+
Ca:AI ratio,	and exchangeable Al). Percentage base saturation was the best
exchangeable	predictor of BAI (nonlinear) and explained 43% of variance.
Al, Ca, Mg, Mn, Multifactorial relationships indicated that tree age and soil
and K (forest	exchangeable Al accounted for 51% and tree age and log of the ratio of
floor and 0 to	base cations (Ca + Mg + K):Al in the soluble (water and acid soluble
15 cm of	wood extracts) fractions accounted for 46% of the variation in sugar
B-horizon)	maple BAI.
June 2018
5-52
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Bondi et al. (2016)
White, Green,
and Adirondack
Mountains
(34 forest sites)
Red-backed NA
salamander
(Plethodon
cinereus)
Soil pH	Studied the relationships between soil pH (organic horizon) and the
(Oa-horizon—on	abundance and health of red-backed salamanders. No associations
16 sites	between soil pH and salamander metrics (abundance, body size, and
A-horizon). Soil	body condition) were found, and the salamanders did not appear to
pH range was	select habitats based on soil pH. The strongest driver of the abundance
2.73-5.54.	of red-backed salamanders was the presence of dusky salamanders
(Desmognathus spp.).
Cai et al. (2017a)
Semiarid
grassland
Inner Mongolia,
China
Grasses
{Agropyron
cristatum,
Stipa
krylovii),
forbs
(Artemisia
frigida,
Potentilla
tanacetifolia,
Potentilla
bifurca)
0, 5, 10, and Soil pH and Al N additions resulted in decreased pH and increased soil available Fe,
15 g N/m2/yr) plant tissue Zn, Mn, and Cu concentrations, with water additions partially counteracting
and two water Cu, Mn, and Fe the impacts of N. Nitrogen additions caused higher foliar Mn, Cu, and
addition	concentrations Zn and lower Fe concentrations, resulting in micronutrient imbalances,
treatments	Similar to the soils, water additions partially offset the impacts of the N
additions on foliar chemistry.
Chen et al. (2013) Grassland
Mongolian
steppe, China
Soil	Seven	Soil pH and Fungal fatty acids were increased by 49% and fungi:bacteria ratio
microbes treatments of S	extractable increased by up to 120% by the H2SO4 additions, relative to the
(bacteria, additions as	cations (Al, Ca, controls. The H2SO4 treatments decreased total and bacterial fatty
fungi, and sulfuric acid (0,	Mg, and NA; 0 acids by up to 47 and 40%, respectively. These responses were
nematodes) 2.76,5.52,	to 15 cm)	attributed to soil pH and Al3+concentrations. High Al3+concentrations
8.28, 11.04,	(51 to 83 mg/kg) were associated with decreased total fatty acids and
13.8, and	decreases in bacterial and increases in fungal fatty acids. Soil
16.56 mol	nematode numbers were initially increased by the H2SO4 treatments
H+/m2) as three	followed by changes in the nematode community. The shifts in the
additions	nematode community were attributed to decreased soil pH and
(2009-2010)	changes in soil moisture.
June 2018
5-53
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Chen etal. (2016)
Alpine	Soil
grasslands and microbes
NA
global
terrestrial
biomes
(bacteria,
fungi)
Soil pH
Studied the relative importance of site, biotic, and climatic factors on
soil microbial communities in two alpine grasslands in the Tibetan
Plateau and across global terrestrial biomes. Microbial communities
were characterized by phospholipid fatty acid (PFLA) and grouped as
total microbial biomass, fungal biomass, arbuscular mycorrhizal fungi
biomass, actinomycete biomass, and fungal:bacterial ratio. In the
Tibetan grasslands, all measures of soil microbial communities were
found to be negatively related to soil pH (soil pH ranged from -6-10)
and positively correlated with soil C:N ratio (C:N ratio range of 4-15).
The fungal:bacterial ratio showed a positive relationship with pH at soil
pHs that ranged from 7.5-10, and negative relationship with soil C:N.
Across the global biomes, total microbial biomass was also found to be
positively correlated with soil C:N and negatively correlated with soil
pH. Soil variables (alone) accounted for 43.4% of the variation in total
microbial biomass, while climatic and biotic variables (alone) only
accounted for 4.5 and 0.2% of variation, respectively.
Cleavitt etal. (2014)
Forest
Hubbard Brook
Experimental
Forest, NH
Sugar
maple
NA
Exchangeable Soil Ca concentration exhibited a 9x change across the study sites and
Ca (top 5 cm of was positively correlated to sugar maple abundance and initial seedling
B-horizon)	densities. However, soil Ca concentration was not a significant
predictor of first-year mortality, nor was it a factor that distinguished
among the three main site types.
Duchesne and
Ouimet (2009)
Forest
Southern
Quebec,
Canada
(426 monitoring
plots)
Sugar
maple and
American
beech
NA
Soil pH,	The basal area of sugar maple in the sapling stratum was positively
exchangeable correlated with soil exchangeable Ca and Mg. Basal area of American
K, Ca, Mg, H, beech in the sapling stratum was negatively correlated with
and Al (upper exchangeable Ca and Mg. However, the basal area of sugar maple in
B-horizon)	the sapling stratum was positively correlated with both the relative
basal area of dead sugar maple and sugar maple in the tree stratum.
The basal area of American beech was also positively correlated with
the relative basal area of American beech in the tree stratum.
June 2018
5-54
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Elias et al. (2009)
Forest
Monongahela
Hardwood
and conifer
National Forest, *ree sPecies
WV (FIA plots)
NA
Soil pH, base The study authors found that hickories were the only species to be in
saturation,	significantly lower numbers on sites with base saturation below 20
exchangeable (A-horizon) and 2.5% (B-horizon). Percentage of dead northern red oak
Ca:AI (A- and was highest on sites with A-horizon Al concentrations above
B-horizons) 43 cmolc/kg of soil. Soil exchangeable Ca:AI and sum of base cations
in the B-horizon were highest in stands that experienced the lowest
species turnover rates (1989-2000). Periodic mean annual volume
increment (whole stand) was positively correlated with A-horizon base
saturation (range of 5 to 77%), Ca concentrations, and exchangeable
Ca:AI (range of 0.17 to 10.2) and B-horizon pH. The exchangeable
Ca:AI of the A-horizon accounted for over 30% of the variation in the
periodic mean annual volume increment.
Gilliam et al.
(2011b)
Forest
Fernow
Experimental
Forest, WV
Soil
microbes
(bacteria
and fungi)
NA
Soil pH (5 cm) Fungi were dominant at the most weathered, low NO3" site, while
Gram-negative bacteria were significantly higher at the less weathered,
moderate and high NO3" sites. Accordingly, the fungi:bacteria ratio
increased in the direction of the low NO3" plots in ordination space.
Correlations between the soil parameters and PLFA results suggest
that low soil pH and NO3" concentrations supported fungal dominance,
although other factors including differences in plant community and
clay and organic matter content may have also influenced the soil
microbial community.
Horslev et al. (2008)
Forest
Pennsylvania,
New York, New
Hampshire, and
Vermont
(86 northern
hardwood
stands)
234 forest
understory
species
NA
The study authors found that a base cation-acid cation nutrient gradient
accounted for 71.9% (in NH and VT) and 63% (in PA and NY) of the
variation in the nonmetric multidimensional scaling ordination analyses
of plant community composition. Sugar maple foliar Mg and Ca had the
strongest association with the base end of the gradient. Exchangeable
Al (in NH and VT) and foliar Mn (in PA and NY) were strongly
associated with the acid end of the gradient. A total of 50 of the
234 understory species were associated with the base end of the base
cation-acid cation nutrient gradient.
June 2018
5-55
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Huang et al. (2017)
Forest
Eastern China
Soil
microbes
(Archaeal
diversity)
NA
Soil pH and Al Studied influence of Al additions on the archaeal diversity in red soils,
concentrations Three Al treatments of 0, 100, or 200 mg Al/kg soil (as AICI3 6H2O); pH
maintained at pretreatment level. Al additions were found to increase
the abundance but decrease the evenness of the Archaea.
Abundances of Crenarchaeota increased while those of Euryarchaeota
decreased in response to higher Al concentrations, suggesting that
Crenarchaeota is more tolerant of Al than is Euryarchaeota
Kunito et al. (2016) Forest
Japan
(27 sites)
Soil
microbes
NA
pH and Al	Evaluated the relationship between soil chemistry (Al concentrations
(soluble and and pH) and soil microbial biomass and enzymes involved with C, N,
exchangeable) and P cycling (p-D-glucosidase, polyphenol oxidase, L-asparaginase,
concentrations acid phophatase). The researchers found that p-D-glucosidase and
in 0-15 cm of polyphenol oxidase were reduced with higher amounts of soluble and
soil	exchangeable Al, acid phosphatase shared an inverse relationship with
soil pH, and L-asparginase increased as pH increased. Microbial
biomass was also found to decrease as organically bound Al and Fe
increased.
Li et al. (2016a)
Natural steppe
ecosystem
Inner Mongolia,
China
Soil	Elevated N Soil pH
microbes 5-15gN/m2/yr (0-15 cm)
(bacteria) as urea app|jed
in May and
June since
2005
Evaluated: (1) the impacts of N and water additions on soil microbial
community (0-15 cm of soil), (2) the linkages between the variation in
belowground bacterial community and aboveground vegetation
community, and (3) relationships between soil/plant factors and soil
microbial community. In general, bacterial alpha diversity was positively
correlated with total N and pH and negatively correlated with soil C:N
ratio and concentrations of NH4 and NO3. Beta diversity of the bacterial
community was significantly correlated with C/N ratio, inorganic N, and
pH. The relative abundances of Proteobacteria, Firmicutes, TM7, and
OD1 increased with N addition gradient. The relative abundances of
Proteobacteria Firmicutes, TM7 and OD1 were positively correlated
with inorganic N and negatively correlated with soil pH.
June 2018
5-56
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Long et al. (2009) Forest	Sugar
Pennsylvania, maple and
New York, New black cherry
Hampshire, and
Vermont
(76 sites)
NA
Soil Ca:AI	Exchangeable Ca, Mg, and pH in upper B-horizon were positively
(threshold of correlated with sugar maple BAI growth in 1987-1996. Generally, in
0.03; upper long-term trends (1937-1996), sugar maple in stands with
B-horizon)	below-threshold amounts of foliar Ca or Mg and above-threshold
amounts of Mn had decreasing BAI trends, while stands with
above-threshold Ca and Mg and below-threshold Mn showed a leveling
off of the BAI. Black cherry consistently showed greater growth on
stands with below-threshold foliar Ca and Mg compared with
above-threshold stands. Black cherry BAI was also greater in stands
with below-threshold Ca:AI molar ratios in the upper B-horizon.
Lucash et al. (2012)
Forest
19 sites in
northeastern
U.S. (New York
and New
Hampshire)
Hardwood NA
and conifer
tree species
Exchangeable
Ca, Mg, and Al
(Oie, Oa 0 to
10 cm, 10 to
30 cm, and
30 cm to top of
C-horizon)
Concentrations of Ca and Mg in foliage were correlated with
exchangeable Ca and Mg concentrations in the upper mineral soil; for
most tree species they were also correlated to acid-extractable Ca and
Mg in the parent material (C-horizon). Foliar Al was insensitive to soil Al
concentrations.
Medeiros et al.
Greenhouse
Red maple
N and P
PH
Studied the influence of pH (simulated acid rain) on leaf, xylem, and
(2016)
study
and white
oak
treatments
(control and low
N/P—90% of
control N/P
levels), pH (4.5
and 6)

hydraulic trait coordination responses of 1-yr old red maple and white
oak seedlings in a greenhouse study. The researchers found
interactions between nutrient levels and pH; low pH reduced the ability
of both species to adjust xylem traits and leaf water relations
(i.e., hydraulic acclimation) in response to changes in nutrient
availability.
McEathron et al.
Forest
Sugar
NA
Soil pH and
Evaluated the relationships between species-specific basal area and
(2013)
Ha-De-Ron-
maple, black

exchangeable
soil and stream water chemistry. Sugar maple basal area was

Dah Wilderness
cherry,

Ca and Al (0 to
positively correlated with mineral soil pH, and yellow birch basal area

Area in
American

10 cm mineral
was positively correlated with mineral soil exchangeable Ca. Sugar

Adirondack
beech, red

soil horizon)
maple basal area was also negatively correlated with stream water

Mountains, NY
maple, and


DOC.

(seven
yellow birch




subwatersheds)




June 2018
5-57
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Mueller et al. (2016)
Sugar maple
dominated
forests
Michigan
Soil
microbes
NA
Soil pH
Studied the relationships between soil bacterial and fungal community
activity and soil pH and C:N gradients in the litter from sugar maple
stands in Michigan. The proportion of active bacteria were found to
increase along the soil pH gradient, but decreased along the soil C:N
gradient. In contrast, no significant correlations were detected for the
fungal community. Similarly, no significant correlations were between
temperature and the fungal or bacterial communities.
Pabian and
Brittinqham (2012)
Forest
Pennsylvania
(14 forest sites)
NA NA
Exchangeable
Ca and pH
(Oa-horizon).
Mean soil
exchangeable
Ca and pH for
the 14 study
sites ranged
from 5.28 to
23.5 meq/100 g
and 3.6 to 5.1,
respectively.
Paqe and Mitchell
Forest
Sugar NA
Exchangeable
(2008)
Adirondack
maple,
Ca,

Mountains, NY
American
exchangeable

(11 sites)
beech,
Ca:AI (forest

American
floor and upper


basswood,
[0 to 10 cm]


and white
mineral soil)


ash

Bird community composition (species richness and species
abundances) varied with soil Ca and pH, with 10 bird species having
the highest abundances in forests with high-Ca soils, and 5 species
having the highest abundances with low-Ca soils. Bird species
associated with low-Ca soils were associated with high densities of
acid-loving mountain laurel (Kalmia latifolia) and five tree species with
basal area explained by low soil pH and Ca, whereas bird species
associated with high-Ca soils were associated with high densities of
saplings and high basal area of acid-sensitive tree species.
Evaluated the relationships between exchangeable soil Ca
concentrations and tree basal area. There were no observed trends
relating total basal area to mineral soil (0 to 10 cm) exchangeable Ca
concentrations; however, the relative basal areas of sugar maple and
American basswood were positively correlated with mineral soil
exchangeable Ca, and relative basal area of American beech was
negatively correlated. Relative basal area of white ash was not
corelated to soil exchangeable Ca.
Perakis et al. (2013)
Forest
Oregon
(coastal range)
Douglas fir NA
(plantations)
None
Nitrate leaching (at 20 and 100 cm) increased, soil pH declined (from
5.8 to 4.2), and exchangeable soil Ca, Mg, and K decreased (10x
declines) along the soil N gradient. Exchangeable Ca and Mg (in both 0
to 20 cm and 0 to 100 cm) and K (0 to 20 cm) declined with increasing
nitrate leaching. Mean soil profile pH declined logarithmically with
nitrate leaching at 20 and 100 cm. The sum of exchangeable Ca, Mg,
and K was positively correlated with soil pH at 20 and 100 cm.
Aboveground tree biomass contained an increasing percentage of total
ecosystem Ca, Mg, and K as soil N increased.
June 2018
5-58
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Pitel and Yanai
(2014)
Forest	Sugar
Massachusetts, maPle
Vermont, and
New York
NA
Soil pH,
exchangeable
Ca:AI, Al, and
cations (Ca, Mg,
K), cation (Ca,
Mg, K, and total)
saturation, and
effective CEC
(A- and upper
B-horizons)
Evaluated the mortality of dominant and codominant sugar maple in
47 stands that had experienced defoliation by native forest tent
caterpillar (Malacosoma disstria). Mortality was found to be highest in
stands with the greatest amount of crown dieback the previous year.
Drought, cold winter temperatures, concave microrelief and soil base
cation availability were also significant predictors of mortality.
Concentrations of exchangeable Ca, Mg, and K in the upper B soil
horizon were inversely correlated with sugar maple mortality, with
exchangeable K having the strongest relationship with mortality. Site
with above-average sugar maple mortality (>3 or 4%) occurred on soils
with low concentrations of exchangeable Ca (0.31 to 0.46 cmolc/kg),
Mg (0.06 to 0.10 cmolc/kg), and K (0.03 to 0.05 cmolc/kg). Stands
defoliated in 2005 that had low Mg saturation (A-horizon) suffered
higher rates of mortality, suggesting an interaction between low base
cations and defoliation events.
Sridevi etal. (2012)
Forest
Hubbard Brook
Experimental
Forest, NH
Soil
microbes
(bacteria)
NA
Soil pH,	Ca additions of 1,000 kg Ca/ha applied in 1999. The bacterial
exchangeable community structure in the Ca treated and nontreated reference soils
cations, Al, Fe, was found to be significantly different, with differences among
P, Mn, Zn,	communities being more pronounced in the mineral soils. Calcium
exchangeable additions resulted in a change in bacterial community composition of
acidity, and 23% in the organic and 22% in the mineral soil horizons. Numbers of
CEC (0 to	detectable taxa in some families were lower in the Ca amended soils,
15 cm)	while some families were higher. Analyses of relationships between soil
chemistry and the bacterial communities indicated that only Ca, pH,
and P were significantly correlated with bacterial community structure.
Stevens et al.
(2010b)
Grassland
U.K. (68 sites)
Grassland
species
6.2 to 36.3 kg
N/ha/yr
Centre for
Ecology and
Hydrology
(CEH) National
Atmospheric
Deposition
Model
Ellenburg R Data from a national survey were used to evaluate species richness of
(reaction-soil 68 U.K. grasslands along an N deposition gradient. The results suggest
pH) and N (soil that soil acidification (instead of eutrophication) was contributing to
nutrient) scores changes in species diversity and composition. Soil acidification may
have led to reduced nutrient availability and increased Al solubility
preventing the "fertilizing" effects of N addition from being apparent.
June 2018
5-59
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Sullivan etal. (2013)
Forest
Adirondack
Mountains, NY
(50 plots in
20 small
watersheds
with sugar
maple
overstory)
Sugar
maple
75 to
112 meq/m2/yr
as N + S
NADP (wet
deposition);
CASTNET (dry
deposition)
2000-2004
(average)
Soil pH,
exchangeable
Ca,
exchangeable
Mg, and base
saturation
(forest floor, A-
and upper B- (0
to 10 cm)
horizons)
The study found that plots with lower soil base saturation did not have
sugar maple regeneration (these same plots also received higher N
and S deposition levels); proportion of sugar maple seedlings dropped
substantially at base saturation levels less than 20%. Canopy vigor was
positively correlated with soil pH and exchangeable Ca and Mg. Mean
growth rates (BAI) were positively correlated with exchangeable Ca
and base saturation at the watershed level.
Tian etal. (2016b)
Temperate
steppe
Inner Mongolia,
China
Grassland
species;
Stipa krylovii
and
Artemesia
frigid a
Elevated N
9 yr of urea
additions at 0,
1, 2, 4, 8, 16,
32, and 64 g
N/m2/yr
Soil pH,
exchangeable
Mn and Al (three
soil depths:
0-10, 10-20,
20-30 cm)
Long-term N additions increased total above-ground plant biomass but
decreased species richness; N additions significantly reduced forb
species richness, while the diversity of grass species was not affected.
Soil chemistry was influenced by the N additions; soil pH was reduced
and concentrations of exchangeable Mn, ferric Fe, and Al were
increased. Foliar concentrations of Mn in both A. frigida and S. krylovii
were increased by N additions. A greenhouse study showed that the
biomass of A. frigida seedling shoots and roots were significantly
reduced with MnCte additions, but the treatments had no effect on S.
krylovii seedlings.
Tu etal. (2016)
Forest types
Soil NA
Soil pH
The soil diazotrophic community structure was found to differ

across North
diazotrophs
(0-10 cm)
significantly across the six forests; lower microbial spatial turnover and

America
(N2 fixing
microbes)

greater community diversity were found in rainforests relative to
temperate forests. In addition, community diversity was strongly
correlated with latitude, mean annual temperature, plant species
richness, and total annual precipitation. Diazotrophic community
diversity was weakly correlated with soil pH and moisture.
Yina etal. (2017)
Grassland
Inner Mongolia,
China
Soil
microbes
(ammonia-
oxidizing
bacteria and
Archaea)
Elevated S
H2SO4 additions
started in 2009;
0, 2.76, 5.52,
8.28, 11.04,
13.80,
16.56 mol H/m2
Soil pH and Study evaluated the impacts of soil pH on the abundance and structure
NH4+-N and of soil ammonia-oxidizing bacteria (AOB) and Archaea (AOA)
NC>3"-N	communities. Decreasing soil pH was found to be correlated with
decreases in AOB abundance (pH range of 5.0-7.3) and increases in
AOA (pH range of 5.3-7.3) abundance, except at the lowest pH (pH
5.0), which negatively influenced AOA abundance. Soil acidification did
not significantly influence AOA or AOB community composition.
June 2018
5-60
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Yuan etal. (2016)
Alpine tundra Soil bacteria Elevated N
Niwot Ridge,	20 yr of N and
Colorado	P additions—0
Rocky	and 20 g
Mountain	N/m2/yr and 0
National Park	and 2 g P/m2/yr
(1993-1995), 0
and 10 g
N/m2/yr and 2 g
P/m2/yr
(1996-1997),
same
application
rates every 2 yr
(1998-2009),
10 g N/m2/yr
was added as
Ca(N03)2 to
address soil
acidification
Soil pH (0-10 Soil bacterial communities differed by plant community type, with
cm)	bacterial alpha diversity being significantly correlated with plant
richness and production of forbs. N additions also influenced soil
bacteria, with bacterial communities treated with N (and N + P) being
significantly different than those in the control and P treatments.
Chloroflexi and Bacteroidetes responded positively to the additions of
N, while Acidobacteria and Verrucomicrobia responded negatively to N.
The N additions resulted in reduced soil pH; from wet to dry systems,
the relative importance of N additions on soil pH increased. Of the soil
variables, pH shared the strongest correlations with plant and bacterial
diversity metrics. Structural equation modeling showed that the indirect
effects (as opposed to direct effects) of N additions—changes in soil
pH and plant communities—were the strongest determinants of soil
bacterial community responses.
Zena etal. (2016)
Temperate
steppe
Inner Mongolia,
China
Soil
microbial
Elevated N
6 yr of NH4NO3
communities additions at 0,
60, 120 and
240 kg N/ha/yr
Soil pH (0-10 Plant community biomass was found to significantly increase, species
cm and 10-20 richness decrease, and N concentrations increase with N additions,
cm)	Seven bacterial phyla (mainly rare phyla) in the 0-10cm soil depth
were significantly changed with N additions, while only 1 phylum in the
10-20-cm depth was affected by the additions of N. Hierarchical
structural equation modeling revealed that changes in bacterial
community composition were due to changes in soil pH and plant
composition, while shifts in bacterial richness were attributed to NH4
concentrations.
June 2018
5-61
DRAFT: Do Not Cite or Quote

-------
Table 5-6 (Continued): Mode of action for acidifying nitrogen and sulfur deposition.
Reference
Ecosystem
Type/Region
Species
N and S
Deposition/
Additions
Soil Indicator
Description
Zhang et al. (2016b)
Temperate
steppe
Inner Mongolia,
China
57 vascular
plant
species
Elevated N
6 yr of NH4NO3
additions at 0,
1, 2, 3, 5, 10,
15, 20, and
50 g N/m2/yr
(2x and 12x per
year)
Soil pH (0-10 N additions, at both frequencies, significantly decreased the number of
cm)	new species gained and increased the number of old species lost.
However, the number of new species gained was lower on the low
frequency of N addition plots compared to the high frequency, while the
number of old species lost was not affected by N addition frequencies.
There was a negative correlation between the cumulative gain of new
species and soil pH, NhU concentrations and community biomass
accumulation, while cumulative loss of old species was positively
correlated with these variables.
Basto et al. (2015b) Grassland
Peak District
National Park,
U.K.
found to be correlated with decreases in total and grass seed
abundance, declines in the persistence of H. pulchrum seed, and
declines in damage to C. rotundifolia seed. Seed germination was not
influenced by pH. In soil with pH higher than 5.6, indirect effects of pH
(through increased fungal pathogens) appeared to decrease the
persistence of the seed of all three grassland species. This study
suggested that: acidic soils are associated with increased seed
persistence; the longevity and size of grassland seed banks decline as
soil pH increases; and that pH indirectly influences seed persistence.
261	NA Soil pH	Studied seed bank and seed germination, viability, and damage
grassland	(gradient with	(through seed burial experiment conducted with Scabiosa columbaria,
species;	soil pH range of	Hypericum pulchrum, and Campanula rotundifolia) along a natural pH
seed bank	3.5 to 6.5)	gradient from acidic to calcareous grasslands. Increasing soil pH was
Al = aluminum; BAI = basal area increment; Ca = calcium; CASTNET = Clean Air Status and Trends Network; Cd = cadmium; CEC = cation exchange capacity; CEH = Centre for
Ecology and Hydrology; cm = centimeter; DOC = dissolved organic compound; Fe = iron; FIA = Forest Inventory and Analysis; g = grams; H+ = hydrogen; ha = hectare;
K = potassium; kg = kilograms; m = meter; meq = milliequivalent; mg = milligrams; Mg = magnesium; Mn = manganese; mol = mole; N = nitrogen; NA = not applicable;
NADP = National Acid Deposition Program; P = phosphorus; PLFA = phospholipid fatty acid: N03" = nitrate; S = sulfur; S042" = sulfate; yr = year; Zn = zinc.
June 2018
5-62
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
APPENDIX 6. TERRESTRIAL ECOSYSTEMS:
NITROGEN ENRICHMENT EFFECTS
ON ECOLOGICAL PROCESSES
Appendix 5 examined the biological effects of terrestrial acidification from nitrogen (N)
and sulfur (S) deposition. This appendix characterizes the biological effects of terrestrial
N enrichment that can be caused by atmospheric N deposition. Following an introduction
(Appendix 6.1). this appendix is composed of five major sections: effects on physiology,
growth, and productivity (Appendix 6.2); changes in biodiversity and community
composition (Appendix 6.3); climate modifications of N enrichment effects
(Appendix 6.4); critical loads (Appendix 6.5); and a summary (Appendix 6.6). The first
two sections begin with an introduction (Appendix 6.2.1 and Appendix 6.3.1) reviewing
the previous causal determination and presenting the current causal determination.
Following this introduction, an overview is presented of the mechanisms operating across
ecosystems to link N enrichment to biological change (Appendix 6.2.2 and
Appendix 6.3.2). Appendix 6.2. Appendix 63, and Appendix 65 are further divided into
subsections based on ecosystem type (e.g., forests, grasslands, etc.) or functional group
(e.g., lichens, trees, herbaceous plants). Finally, Appendix 6J> provides a summary of the
new information generated since the 2008 ISA for Oxides of Nitrogen and
Sulfur—Ecological Criteria (hereafter referred to as the "2008 ISA").
6.1. Introduction
Nitrogen (N) is a key element required by all organisms in order to build amino acids and
nucleic acids, the basic biochemical subunits needed to synthesize the proteins, enzymes,
RNA, and DNA sustaining all biological processes. By the second half of the 19th
century, before these biochemical pathways had been identified, it was already
understood that N was a component of plant and animal tissues and a supply of N was
essential for plant growth (Galloway et al.. 2004). Indeed, the ability of added N to
stimulate plant growth had been recognized by science (and commerce) for over a
century (Galloway et al.. 2004; Galloway and Cowling. 2002) prior to the 2008 ISA. By
2008, it was already clear that N availability broadly limited productivity in terrestrial
ecosystems.
By 2008, it was also clear N availability could alter the biodiversity of terrestrial
ecosystems. Broadly, the effects ofN deposition on the diversity of terrestrial ecosystems
stem from four mechanisms [sensu (Bohhink et al.. 2010)1: (1) eutrophication,
(2) acidification, (3) direct toxicity and damage, and (4) increased susceptibility to
June 2018
6-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
secondary stress. Multiple mechanisms likely operate simultaneously to alter diversity
within a particular community or ecosystem. As described in Appendix 5. N deposition
may lead to soil acidification, which can have negative effects on plants and
microorganisms. Direct toxicity and damage from N deposition often comes from an
accumulation of NH44" in soils and plant tissues. Increased susceptibility to secondary
stressors includes greater impacts of pathogens (Bobbink et al.. 2010) and shifts in
herbivory as a result of altered tissue chemistry(Throop and Lerdau. 2004). Among these
four major mechanisms, eutrophication is perhaps the most complex because it can
change the physiology of individual organisms, alter the relative growth and abundance
of species, transform relationships between species, and indirectly affect the availability
of other essential resources such as light, water, and nutrients KHauticr et al.. 2009; Clark
et al.. 2007; Suding et al.. 2005); see also Appendix 41. These can lead to biodiversity
shifts, including community compositional changes, the loss of species, and decline in
species richness.
Since the 2008 ISA, the effects of N deposition on terrestrial ecological processes and
biodiversity have continued to be widely studied. With the increasing volume of research,
a number of new studies have been conducted using meta-analysis to synthesize
published observations or using continental or global data sets to understand broad-scale
patterns. New studies have provided a more detailed understanding of how N influences
terrestrial ecosystem growth and productivity; community composition and biodiversity;
and sensitive organisms and ecosystems. Further, a large body of work has been
published on critical loads (CLs) for N since 2008. Together with the information
available in the 2008 ISA, this body of evidence is sufficient to infer a causal
relationship between N deposition and (1) the alteration of the physiology and
growth of terrestrial organisms and the productivity of terrestrial ecosystems; and
(2) the alteration of species richness, community composition, and biodiversity in
terrestrial ecosystems.
6.2. Linking Nitrogen Deposition to Changes in Physiology,
Growth, and Productivity in Terrestrial Ecosystems
6.2.1. Introduction
In the 2008 ISA, evidence was sufficient to infer a causal relationship between N
deposition and the alteration of the terrestrial carbon (C) and N biogeochemical cycles.
These effects included not only changes in soil C and N pools and fluxes (described in
Appendix 4). but also significant alterations of plant and microbial growth and
June 2018
6-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
physiology. The 2008 ISA built upon the conclusions of the 1993 Oxides of Nitrogen Air
Quality Criteria Document. By 1993, a series of hypotheses regarding the effects of
chronic N deposition on northern temperate forests had already been developed from
early field observations (Aberetal.. 1989). Briefly, N deposition was expected to
increase tree growth in most forests, but once plant demand for N had been satisfied
(biological and physical sinks had reached "N saturation"), further N deposition could
lead to nitrification, soil cation leaching, acidification, nutrient deficiencies, and
decreased tree growth (Aberetal.. 1989) (described in Appendix 4 and Appendix 5). A
revised form of these N saturation hypotheses (Aberetal.. 1998) provided much of the
conceptual foundation in the 2008 ISA for understanding how N deposition influenced
plant physiology, growth, and ecosystem productivity.
The effects of N deposition on terrestrial ecological processes have been widely studied
since 2008. Research on N deposition has continued in North America and Europe, with
other areas receiving less attention [e.g., temperate forests in Asia and the Southern
Hemisphere (Gilliam. 2016)1. A significant new body of research has developed in the
boreal, arid, and subtropical ecosystems of Asia, particularly in China [e.g., (Zhang et al..
2015e; Du et al.. 2014a; Du et al.. 2014b; Du and Fang. 2014; Sun et al.. 2014)1. New
meta-analyses have provided a more detailed understanding of how added N affects
productivity responses in different biomes (LeBauer and Treseder. 2008). growth among
plant functional types (Xia and Wan. 2008). growth of individual plant parts (Xia and
Wan. 2008). root growth (Li et al.. 2015). and ecosystem C storage (Liu and Greaver.
2009). while other broad-scale analyses have examined changes in plant N concentrations
(Xia and Wan. 2008). microbial biomass (Treseder. 2008). and belowground C cycling
(Liu and Greaver. 2010). This expanded body of research has created a better
understanding of how N influences processes at molecular to global scales. Together with
the information available in the 2008 ISA, this body of evidence is sufficient to infer a
causal relationship between N deposition and the alteration of the physiology and growth
of terrestrial organisms and the productivity of terrestrial ecosystems.
6.2.2. Mechanisms Operating across Terrestrial Ecosystems
The 2008 ISA evaluated a large number of studies assessing how N deposition has
changed terrestrial C cycling and found an array of ecological responses. The most
extensive evidence of the effect of N deposition on C cycling was available for forests in
North America and Europe. In experimental N addition studies, moderate to high
additions ofN led to either no significant change in tree growth rates or transient growth
increases (generally at deposition rates lower than 10 kg N/ha/yr), followed by increased
mortality, especially at higher rates of N addition. This group of studies showed
June 2018
6-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
coherence in effects and indicated the presence of a biological gradient in responses with
increasing N deposition.
Some of the broad-scale changes caused by N additions to terrestrial ecosystems have
been quantified by meta-analysis, a data synthesis tool that started to become common in
ecological research beginning in the late 1990s (Gurevitch et al.. 2001). Using this tool to
understand the effects of N additions on herbaceous plant communities, (Gough etal..
2000) found that N additions stimulated aboveground plant productivity by an average of
53%. In a broader synthesis, (Elseretal.. 2007) conducted a meta-analysis of N addition
effects on plant community productivity and observed average increases in productivity
of 20-30% in grasslands, forests, tundra, and wetlands. In addition to changes in plant
growth, (Koricheva et al.. 1998) identified via meta-analysis that added N could alter the
chemistry of plant tissues, including increasing tissue concentrations ofN and free amino
acids, while decreasing concentrations of starch and C based secondary compounds
(important for defense against herbivores), including phenylpropanoid compounds. In a
meta-analysis of the belowground effects of N, Treseder (2004) found N additions
decreased both the abundance of mycorrhizal fungi and the percentage of plant roots
colonized by mycorrhizal fungi. In addition, multiple lines of evidence showed that N
deposition increases the performance of insect herbivores, and potentially, insect
populations (Throop and Lerdau. 2004). In a synthesis of 500 observations of the effect
of N on litter decomposition rates, Knorretal. (2005) found that added N stimulated
decomposition at sites with low rates of ambient N deposition (<5 kg N/ha/yr), but
slowed decomposition at sites with moderate rates of N deposition [5-10 kg N/ha/yr;
(Knorr etal.. 2005)1 and that N additions at rates from 2 to 20 times ambient N
deposition inhibited decomposition by 8 to 16%.
These changes can cause a cascade of ecological consequences. For instance, the
inhibition of decomposition can increase soil C content, but the accumulation of plant
litter on the soil surface can inhibit the establishment of some plant species (Patterson et
al.. 2012; Cleavitt et al.. 201 lb; Xiong and Nilsson. 1999; Facelli and Pickett. 1991). In
the western U.S., Fenn et al. (2003a) suggested greater plant growth caused by N
deposition could increase plant litter accumulation, in turn increasing the susceptibility of
forests and other wildlands to severe wildfires. Nitrogen deposition also affects the
patterns of C allocation because most of the additional plant growth occurs aboveground.
This increases the shoot-to-root ratio, which can be detrimental to plants because of
decreased resistance to environmental stressors, such as drought and wind (Braun et al..
2003; Krupa. 2003; Minnich et al.. 1995; Fangmeier et al.. 1994).Thus. it was recognized
in the 2008 ISA that the effects of N deposition on biological and ecological processes in
terrestrial ecosystems were pervasive, complex, and difficult to fully understand and
predict.
6-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Although evidence showed plant productivity responses to N deposition as of the 2008
ISA, the extent to which N increased forest C content in C budget measurements
(quantified input and output of C to the ecosystem) was uncertain. A prominently
reported estimate of >600 kg C sequestered per kg of N deposited to forests, which was
based on a regional gradient technique (Magnani et aL 2007). was widely criticized.
Subsequent reassessments of these data suggested that forest C sequestration was an
order of magnitude lower (Sutton et al.. 2008). Liu and Greaver (2009) found in a
meta-analysis that N additions ranging from 25.5 to 200 kg N/ha/yr increased forest
ecosystem C content (+6%).
In addition, fewer studies had examined the effects of N deposition on the eutrophication
of nonforested ecosystems. Mack et al. (2004) examined C and N pools in a long-term N
addition experiment at the Arctic Long-Term Ecological Research site near Toolik Lake,
AK. Plant growth increased as a result of N additions, but the N additions enhanced
decomposition of belowground C pools in deep soil layers more than it increased primary
production, leading to a substantial net loss of C from this ecosystem. More broadly, Liu
and Greaver (2009) conducted a meta-analysis of 16 observations from nine publications
to evaluate the relationship between N addition (16 to 320 kg N/ha/yr) and C
sequestration in nonforest ecosystems and did not find a significant effect on net
ecosystem exchange (NEE, kg C/ha/yr).
Xia and Wan (2008) identified nearly 1,600 observations of plant biomass growth in
response to N additions, excluding agricultural and horticultural species (Figure 6-1 A).
Overall, plant biomass increased by 54%, with a larger gain in seeded plants (+55%) than
spore plants (+21%; not shown). Among the seeded plants, grasses showed the largest
biomass response, followed by trees, forbs, and shrubs. Annual herbs (+92%) showed a
stronger response than perennial herbs (+56%). Overall, biomass responses to N
increased linearly with mean annual precipitation (MAP). The studies included in the
analysis had addition rates ranging from 10 kg N/ha/yr to 1,000 kg N/ha/yr. When the N
additions were divided into low (<100 kg N/ha/yr) and high (>100 kg N/ha/yr) groups,
there were only a few differences among functional groups. The response of woody
species was lower than for herbaceous species at the low addition rate (+25 vs. +51%),
but the groups had similar responses at the high N addition rate. Among herbaceous
plants, both legumes and nonlegumes had positive responses to low rates of N additions
(19 and 23%, respectively), while the high rate of N additions increased nonlegume
biomass (+43%), but did not significantly affect legume growth. The authors also
examined how these responses varied in response to the rate of N additions. The N
addition studies were divided into 50 kg N/ha/yr increments from 0 to 300 kg N/ha/yr,
but aside from increasingly positive growth responses for trees and woody plants with
greater N additions, there were few apparent trends, and sample sizes were inconsistent.
6-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
In their analysis, Xia and Wan (2008) separated plant growth data into aboveground
(887 observations) and belowground (340 observations) components. Across almost all
functional groups, there were greater increases in biomass in aboveground than
belowground components (Figure 6-1 A) (see also Appendix 4). In a subsequent
meta-analysis of plant C pools, Luetal. (2011b) also found that N additions to plants in
nonagricultural systems stimulated shoot C by 28.5% (n = 146) and root C by 20%
(n = 77). These results indirectly suggest that plant C allocation changes in response to N
addition. These changes were also shown by a more direct analysis of a much smaller
meta-analysis data set (n = 15) in which Li et al. (2015) observed a significant 11%
decrease in the root:shoot biomass ratio. Together, these results provide further support to
the conclusion in the 2008 ISA that N additions can alter plant C allocation and result in
elevated shoot: root ratios. Further, Luetal. (2011b) analyzed the response of some
individual plant parts to added N and observed a 1.6% increase in leaf C mass.
Belowground processes are important components of terrestrial ecosystems. As detailed
in Appendix 5. soils are often the largest ecosystem pools of both C and N; soil
respiration can be the largest ecosystem efflux of CO2; and soils can be important sources
and sinks for the greenhouse gases N2O and CH4. Biologically, the large C fluxes plants
allocate belowground for root exudation, the growth and maintenance of roots, and the
support of mycorrhizal fungi help support complex belowground foodwebs. Less
information is available about belowground responses to N additions than aboveground
responses. However, Li et al. (2015) recently conducted a meta-analysis of how root traits
respond to N additions. Notably, Li et al. (2015) found that although total root biomass
increased (+20%; n = 74), fine root biomass declined (-13%; Figure 6-2B). This
discrepancy is apparently accounted for by an increase in the biomass of coarse roots
(+57%; n = 7), which are typically structural and conductive tissues. Morphological traits
of fine roots, such as length (n = 25), diameter (n = 10), and specific root length (n = 22),
were unaffected by N additions. Fine root turnover rate was 21% higher (n = 12), but fine
root production was not significantly affected by the N additions [n = 22; (Li et al..
2015)1.
Aboveground net primary productivity (NPP), which measures aboveground plant growth
at the community scale, was less responsive on average (+29%) than individual plant
growth to N additions I(LeBauer and Treseder. 2008); Figure 6-IB and Figure 6-2A1.
Consistent with the plant functional group analysis conducted by Xia and Wan (2008).
LeBauer and Treseder (2008) observed a large stimulation of aboveground NPP in
grass-dominated ecosystems such as temperate grasslands (+53%), tropical grasslands
(+26%), and tundra (+35%), whereas NPP increased only 19% in temperate forests.
There were also significant NPP increases in tropical forests and wetlands. Added N did
6-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
not significantly increase NPP in deserts, but the analysis only included three
observations.
More recently, Tian et al. (2016a) synthesized aboveground NPP data from studies that
had experimentally added N at more than one rate in order to examine how plant growth
responses per unit N changed with increasing levels of N. Their metric, termed "N
response efficiency" (100% x [ANPPtreatment - ANPPControi]/ANPPControi/N addition rate) did
not significantly differ across the three ecosystem types (wetlands, forests, grasslands),
with an average increase of 3 to 4% in NPP per g of added N. The N response efficiency
decreased with N addition rates above 50 kg N/ha/yr, consistent with a saturating
response to N.
Liu and Greaver (2009) had smaller data sets, with 16 observations of net ecosystem
exchange (NEE; all in nonforested systems) and 17 observations of ecosystem C content
(Figure 6-2). There were no significant effects of added N on NEE overall or within the
individual biomes included in the analysis (grasslands, wetlands, tundra). However,
added N did increase forest ecosystem C content by 6%. Belowground NPP responses
have not been synthesized, at least in part, because data are lacking (LeBauer and
Treseder. 2008).
In terms of altered physiology, there is also widespread evidence that N additions
increase plant tissue N concentrations. The meta-analysis by Xia and Wan (2008)
included changes in plant N concentrations overall, aboveground, and belowground.
Overall, N additions increased plant N concentrations by an average of 28.5% (Xia and
Wan. 2008). Although there was significant variation among plant functional groups, the
average increase in tissue N concentration among seeded plants was much more similar
across functional groups than the biomass growth response, ranging only from +24 to
+35% once legumes (+7%) and the two broader functional groups containing legumes
(forbs, +14%; perennial herbs: +22%) were excluded. Likewise, belowground and
aboveground plant tissues had a similar change in N concentration, in contrast to the
varying biomass growth responses to N (Xia and Wan. 2008). Li et al. (2015) also
observed that N additions significantly increased root N concentrations in a root trait
meta-analysis (+18%; n = 84). Because of these increased N concentrations, N additions
can significantly decrease plant C:N ratios (Yue et al.. 2017).
6-7
DRAFT: Do Not Cite or Quote

-------
NPP = net primary productivity.
Notes: (A) Mean change in plant biomass growth in response to N additions redrawn from Xia and Wan (2008). with blue bars
representing the overall response, orange bars representing aboveground growth, and grey bars representing belowground growth.
Vertical grey lines identify different ways the data were parsed (i.e., all seeded plants; woody vs. herbaceous plants, etc.). (B) Mean
change in aboveground net primary productivity, from LeBauer and Treseder (2008). Error bars represent the 95% confidence
interval. Numbers above the error bars indicate the number of observations included in the analysis.
Figure 6-1 Effects of nitrogen additions on plant growth and net primary
productivity.
June 2018
6-8
DRAFT: Do Not Cite or Quote

-------
B
60
50
40
?30
ai
.Si 20
on
4—«
o
»2i 10
0
-10
-20
-30
60
50
40
I30
Ol
20
<-o
*—>
u
Ł 10
0
-10
-20
-30
Total
Pool
Below
Ground
Pools
37
Above
Ground
Biomass
Net Ecosystem
Flux Inputand
Output
Fluxes
53
rrn
Soil
Organic SoilC DOC
Horizon
57
31
Mycorrhiza
Microbial
C
Total
Plant
Fluxes
126
16
No No
AnalysisAnatysis
No No
AnalysisAnatysis
Below
Ground
Fluxes
500
No
GPP R,
ecosystem
Above Below RK01Ą„trr.
Ground Ground
NPP NPP
i?
Analysis ^7
autot^p^
Decomposition
SO-:
hetr-otroph
NEE
DOC = dissolved organic carbon, NEE = net ecosystem exchange, GPP = gross primary production, Recosystem = ecosystem
respiration, NPP = net primary production; Recosystem autotroph = plant respiration, Rsoj = soil respiration, Rson heterotroph = heterotroph soil
respiration.
Notes: mean effect sizes from meta-analyses of N addition experiments on ecosystem, plant, and soil pools ([a] top panel) and
fluxes ([b] bottom panel). Error bars represent 95% confidence intervals. Numbers above the bars are the sample sizes for each
analysis. Letter with each bar denote the data source. No Analysis denotes a pool or flux that has not yet been meta-analyzed.
Data sources: (A) Liu and Greaver (2009). (B) Xia and Wan (2008). (C) Lu et al. (2011b): (D) Li et al. (2015); (E) Liu and Greaver
(2010). (F) Treseder (2004): (G) LeBauer and Treseder (2008): (H) Knorr et al (2005).
Figure 6-2
Effects of added nitrogen on ecosystem carbon pools and fluxes.
June 2018
6-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
The 2008 ISA linked enhanced terrestrial productivity to increases in photosynthesis and
gross primary productivity. Gross primary productivity (GPP) can increase as a result of
either a greater amount of photosynthetic tissue (more light absorption) or a higher rate of
photosynthesis per leaf. There is a strong mechanistic and conceptual link between
greater foliar N concentrations, higher leaf concentrations of the C assimilating enzyme
rubisco (Evans. 1989). and greater maximum rates of leaf-level photosynthesis in
vascular plants (Wright et al.. 2004). Consequently, the increase in foliar N caused by N
additions has been linked to increases in the leaf-level rate of photosynthesis for decades
[e.g., (Teskev et al.. 1994) and references therein]. Alternately, evidence cited in the 2008
ISA suggests that much of the increased foliar N observed when N is added may be
physiologically inactive because it manifests as an increase in storage compounds such as
free amino acids (Bauer et al.. 2004). While increases in photosynthesis in response to N
additions have been observed in trees, grasses, and shrubs, these increases have not been
consistent [e.g., (Pivovaroff et al.. 2016; Talhelm et al.. 2011; Elvir et al.. 2006; Chen et
al.. 2005b; Newman et al.. 2003; Laitha and Whitford. 1989; Gulmon and Chu. 1981)1
and there does not appear to be a meta-analysis or other synthesis on the response of
leaf-level photosynthesis or GPP (Figure 6-2B) to N additions. Similarly, there is a strong
fundamental relationship between tissue N concentration and respiration rates in plants
(Reich et al.. 2008; Reich et al.. 2006; Ryan et al.. 1996). such that this relationship is
used to model respiration rates [e.g., (Hanson et al.. 2004; Amthor. 2000)1. However,
there is evidence that this relationship can breakdown in N addition studies [e.g., (Burton
et al.. 2012; Drake et al.. 2008; Schaberg et al.. 1997)1. and there are currently no broad
analyses on the effects of N additions on ecosystem or plant-scale autotrophic respiration
(Figure 6-2B).
The 2008 ISA noted some observations of decreased microbial biomass as a result of
added N, particularly for mycorrhizal fungi lYTreseder. 2004); Figure 6-2B1. Since 2008,
however, it has become increasingly clear that N deposition can greatly impact microbial
communities, often including a decrease in microbial biomass. Meta-analyses conducted
across all ecosystem types have found that N additions can decrease microbial biomass
(Treseder. 2008). microbial biomass C (Liu and Greaver. 2010). and microbial biomass N
(Lu et al.. 2011b) and that the effects of added N on microbial biomass increase with the
duration of N additions and the total amount of N added (Treseder. 2008). These changes
are noteworthy given the diverse role of microorganisms in nutrient cycling, greenhouse
gas fluxes, and other ecosystem services. Although N additions frequently decrease
microbial biomass, the results are not always consistent (Liu and Greaver. 2010;
Treseder. 2008). The effects of N were not significant at the level of individual microbial
domains (bacteria, fungi), although there were trends toward increasing negative effects
of N additions on fungal biomass as duration and cumulative N load increased (Treseder.
2008). In a meta-analysis, Carey et al. (2016) observed that N additions increased the
6-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
abundance of ammonia-oxidizing bacteria across nonagricultural ecosystems, but did not
affect the abundance of ammonia-oxidizing archaea. Recently, Li et al. (2015) provided a
new meta-analysis ofN addition effects on mycorrhizal colonization of fine root tips and
observed an overall 17% decline (n = 21). While this reinforces the conclusions of an
earlier meta-analysis by Treseder (2004). there are important qualifiers to this finding: the
sample only included observations from forests and grasslands in temperate and
subtropical climates; significant decreases were only observed in temperate and forested
ecosystems, after more than 3 years of N additions, or at N addition rates
>100 kg N/ha/yr (Li et al.. 2015).
Changes in mycorrhizae in turn can affect plant productivity. Mycorrhizal fungi provide
benefits to plants, yet also carry costs. In exchange for nutrients and water from the
fungus, the plant provides C from photosynthesis (Hogbcrg et al.. 2010; Rillig. 2004). In
many cases, this tradeoff works to the plant's advantage. For instance, Van Per Heiiden
et al. (1998) observed higher biomass for most of the individual plant species in their
study when inoculated with arbuscular mycorrhizae versus when mycorrhizae were
absent. They suggested the mycorrhizal acquisition of a limiting nutrient, in this case
phosphorus (P), for the plant in exchange for C explained this finding. However, under
conditions in which the nutrient limitation of the plant is relieved (e.g., high N
availability), mycorrhizae may no longer be as needed by the plant for nutrient
acquisition, yet concomitantly still impose a C cost. In this case, the tradeoff is no longer
as beneficial to the plant. With increased N availability, plants that make fewer
mycorrhizal associations can benefit both from the physiological advantages
accompanying an alleviation of N limitation (e.g., increases in photosynthesis, changes in
plant chemistry) and from incurring a lower C cost for mycorrhizae (Johnson et al..
2008). Indeed, shifts in C allocation away from mycorrhizae may be a mechanism to
support increased plant productivity in situations where plants do not show gains in
photosynthesis [e.g., (Talhelm et al.. 2011)1.
There are numerous mechanisms through which N deposition could impact microbial
biomass, including changes in soil chemistry, changes in the rates of aboveground and
belowground plant C inputs (including litter production, root exudates, and C supplied to
mycorrhizal fungi), and changes in plant tissue chemistry (Treseder. 2008). Nitrogen
additions can change the chemistry of litter inputs by altering the tissue composition of
plant species (Throop and Lerdau. 2004) or by changing the composition of plant
communities contributing to the ecosystem flux of plant litter [e.g., (Manning et al..
2008)1. Within a particular species, N additions can cause changes in both of the
dominant forms of plant litter: leaf litter and fine roots [e.g., (Xia etal.. 2015)1. These
changes can be relatively direct, such as increases in tissue concentrations of inorganic
and organic forms of N [e.g., (Bauer et al.. 2004; Koricheva et al.. 1998)1. While leaf
6-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
litter N concentrations are not as widely measured as green leaf N, increases in leaf litter
N concentrations due to added N are also likely to be widespread because increased N
availability often either decreases or has no effect on the fraction of N plants absorb from
leaves during senescence (Acrts. 1996). Other changes in litter chemistry, such as
increases in condensed tannins, soluble phenolics, and nonstructural carbohydrates (Xia
ct al.. 2015). likely involve more complex biological mechanisms. In a meta-analysis of
plant biochemistry data, Liu et al. (2016a) observed different effects of added N on trees
and herbs. For the live tissues of trees, N additions decreased concentrations of lignin,
cellulose, and nonstructural carbohydrates, and increased the concentration of protein. In
the live tissues of herbs, added N decreased the concentrations of nonstructural
carbohydrates and hemicellulose, but increased lignin, cellulose, and protein
concentrations. These changes have implications for herbivores and detritivores, as well
plant growth dynamics.
In a litter decomposition meta-analysis, Knorr et al. (2005) observed that the effects of
added N also differed based on the length of the study and the initial litter lignin content.
Added N stimulated decomposition in studies lasting less than 2 years, but reduced
decomposition in studies lasting more than 2 years. High-lignin litter decomposed more
slowly than low-lignin litter. Because lignin concentrations tend to increase during litter
decomposition, both of these results were considered to be consistent with earlier
evidence that increased N supplies stimulate microbial activity responsible for the
decomposition of labile litter constituents, while suppressing the microbial production of
extracellular enzymes responsible for the degradation of lignin (Knorr etal.. 2005) (see
also Appendix 4). In a meta-analysis of biochemistry data for decomposing litter, Liu et
al. (2016a) observed that N additions increased the concentration of lignin in
decomposing litter for herbs and shrubs, but did not have a significant effect on
decomposing tree litter. In comparison, N additions decreased cellulose concentrations in
decomposing tree litter, but did not significantly affect cellulose concentrations in
decomposing herb litter. In another meta-analysis, Liu and Greaver (2010) found that
although N additions increased aboveground litter production (+20%; n = 37; not shown)
across all biomes and N addition rates, there were no significant overall changes in either
total soil respiration (Figure 6-2B) or heterotrophic soil respiration (Figure 6-2B). In fact,
both forms of soil respiration tended to decline. Currently, no similar cross-biome
analysis has been conducted for changes in autotrophic soil respiration (respiration from
roots and mycorrhizal fungi), but Janssens et al. (2010) noted in a meta-analysis that
autotrophic respiration was suppressed by N additions in forests.
Because some biogeochemical processes involve specific chemical forms of N
(e.g., denitrification, ammonium toxicity) (see Appendix 4). there is the potential that
biological responses to N deposition (or N addition) could depend on whether the
6-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
dominant form of deposited N is oxidized (NOy) or reduced (NHx). A number of studies
have specifically addressed this issue, either by conducting experiments directly testing
additions of different forms of N or indirectly through syntheses comparing the effects of
NOy and NHx in different experiments. Different responses to individual forms of N have
been observed for some endpoints, such as increases in dissolved organic C, decreases in
ecosystem N retention, increases in soil N2O emissions, plant growth, and defense against
pathogens (Mur et al.. 2016; Yue et al.. 2016; Templer et al.. 2012; Verhoeven et al..
2011; Liu and Greaver. 2010. 2009) (see also Table 4-13). By contrast, other studies have
failed to observe a difference between the effects of N forms. One direct test occurred in
the Front Range of the Rocky Mountains in Colorado, where Ramirez et al. (2010a)
investigated whether soil microbes respond differently to additions of NH4+ versus NO3
and found that the total amount of N added was correlated with a decrease in soil
respiration, not the form of the N (NFU+ vs. NO3 ). A more comprehensive understanding
is available by reviewing the results of meta-analyses comparing the responses of N
addition experiments conducted with different forms of N (Table 6-1). With notable
exceptions, most often differences in the effect of the form of N were not observed in
these meta-analyses. Moreover, studies finding differences tended to occur where sample
sizes were small [e.g., belowground C pools in (Yue et al.. 2016)1.
6-13
DRAFT: Do Not Cite or Quote

-------
Table 6-1 The effects of different forms of inorganic nitrogen on biological
endpoints according to meta-analyses.
See Table 4-13 for the effects of different forms of inorganic nitrogen on
biogeochemical processes and indicators.
Reference
Endpoint
Effect of NOy vs. NHx Forms
Yue et al. (2016)
Aboveground plant C pool
Not significant
LeBauer and Treseder (2008)
Aboveground plant productivity
Not significant
Yue et al. (2016)
Aboveground plant productivity
Not significant
Treseder (2008)
Bacteria biomass
Not significant
Yue et al. (2016)
Belowground plant C pool
Increase with NhU"1"
Liu and Greaver (2009)
Ecosystem C content
Not significant
Liu and Greaver (2010)
Fine root litter production
Not significant
Treseder (2008)
Fungal biomass
Not significant
Liu and Greaver (2010)
Leaf litter production
Insufficient data
Yue et al. (2016)
Leaf litter production
Insufficient data
Knorr et al. (2005)
Litter decomposition
Not significant
Yue et al. (2016)
Litter decomposition
Not significant
Treseder (2008)
Microbial biomass
Not significant
Liu and Greaver (2010)
Microbial biomass C
Decrease with NHV
Yue et al. (2016)
Microbial biomass C
Not significant
Treseder (2004)
Mycorrhizal abundance
Not significant
C = carbon; 15N = tracer isotope of nitrogen; N20 = nitrous oxide; NH4+ = ammonium; NH4N03 = ammonium nitrate; NHX = sum of
reduced forms of N; N03" = nitrate; NOY = the sum of oxidized forms of nitrogen. Notes: References ordered by endpoint. Only
statistically significant differences between the effects of forms of N listed as increases or decreases.
1	Before the 2008 ISA, neither terrestrial N cycling nor anthropogenic N deposition had
2	been widely incorporated into Earth systems models (ESMs) used to understand and
3	forecast global climate and biogeochemical cycling. Thornton et al. (2007) made the first
4	effort to understand how both coupled C and N cycling and anthropogenic N deposition
5	would impact ESM predictions for terrestrial C uptake by inserting a new land
6	biogeochemistry model into a coupled climate system model. In the resulting model
7	output, N limitation greatly decreased the amount of terrestrial C uptake predicted from
June 2018
6-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
future increases in atmospheric CO2 and decreased the sensitivity of terrestrial C
sequestration to increases in temperature and precipitation (Thornton et al.. 2007).
Including N deposition within the model directly increased terrestrial C uptake and also
indirectly increased terrestrial C uptake by removing some of the N limitation predicted
to occur under increased atmospheric CO2 (Thornton et al.. 2007). Subsequently, efforts
to model biogeochemical processes at regional and global scales have expanded (Thomas
et al.. 2015; Zaehle and Dalmonech. 2011). Although the integration of terrestrial N
cycling into ESMs remains relatively weak, more ESMs are incorporating the coupling of
terrestrial C and N cycling as overall model development and sophistication advances
(Thomas et al.. 2015; Arora et al.. 2013; Zaehle and Dalmonech. 2011; Bonan and Levis.
2010; Gerber et al.. 2010). Consistent with the findings of Thornton et al. (2007).
inclusion of coupled C-N cycling in ESMs has two primary effects: decreasing the
stimulatory effects of elevated atmospheric CO2 on terrestrial productivity and decreasing
the sensitivity of terrestrial C sequestration to climate warming because increased soil N
mineralization stimulates plant productivity (Arora et al.. 2013; Zaehle and Dalmonech.
2011; Zhang et al.. 201 lb; Arneth et al.. 2010; Bonan and Levis. 2010; Gerber et al..
2010; Yang et al.. 2010; Zaehle et al.. 2010; Thornton et al.. 2009). However, the ESMs
that do include basic terrestrial C-N coupling lack more recently identified interactions
such as plant organic N uptake, soil priming (root exudation), and the suppression of litter
decomposition at high soil N availability, all of which could increase terrestrial C uptake
to varying extents (Thomas et al.. 2015; Zaehle and Dalmonech. 2011). In addition, the
interactions between and the net effects of N, precipitation, and temperature on
ecosystem C response in soils are unknown in many cases (see Appendix 13.1.2.1).
In the few ESMs that have directly included the effects of N deposition, the additional N
increased terrestrial C uptake and increased the extent to which elevated atmospheric CO2
stimulates terrestrial C uptake (Devaraiu et al.. 2016; Bonan and Levis. 2010; Yang et al..
2010). In addition, N deposition in the Northern Hemisphere (particularly in the U.S.)
apparently compounded the increase in C sequestration caused by the regrowth of
secondary forests during the late 20th century following agricultural abandonment and
timber harvest (Gerber et al.. 2013; Yang et al.. 2010). However, ESMs that do not
include potential N saturation may overestimate the effect of N deposition on terrestrial C
uptake in regions experiencing high N deposition rates (Lu et al.. 2016). In addition, it
should be noted that although N deposition and the overall anthropogenic production of
reactive N increases terrestrial C sequestration, they are not the only influence of
anthropogenic N on global climate. It is difficult to quantify the overall climate impact of
anthropogenic N [e.g., (Pinderetal.. 2013)1 because reactive N can change the planetary
albedo by enhancing aerosol formation, stimulate the production of biogenic greenhouse
gases, alter the production and destruction of methane and tropospheric ozone in the
atmosphere, and indirectly decrease terrestrial productivity due to the phytotoxic effects
6-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
of ozone (Lu and Tian. 2013; Pinder et al.. 2013; Pinder et al.. 2012; Arneth et al.. 2010;
Zaehle et al.. 2010). Many of these effects are discussed further in the atmospheric
chemistry and terrestrial biogeochemistry portions of this ISA (see Appendix 2 and
Appendix 4).
In addition to the global-scale analyses conducted using ESMs, biogeochemical process
models have been used to assess the impact of N deposition on terrestrial productivity
and C sequestration at national and regional scales. Tian et al. (2012) used the Dynamic
Land Ecosystem Model (DLEM) to model the influence of climate, tropospheric ozone,
fertilizer use, land use/land cover change, atmospheric CO2, and N deposition on
terrestrial C storage over the 20th century in the southeastern U.S. (encompassing Texas
and Oklahoma to Florida and Virginia). Terrestrial C storage increased from 1951-2007,
with the model identifying atmospheric CO2 and N deposition as the environmental
factors responsible for the increase in C sequestration (Tian et al.. 2012). In China, a
series of papers has been published using biogeochemical process models to identify how
this same set of environmental factors influenced terrestrial C cycling in that country.
Rates of N deposition in China grew rapidly during the late 20th century and were
considerably higher than in the U.S. by the early 21st century, with nationwide average
rates of approximately 20 kg N/ha/yr and rates in southeast China averaging 35-40 kg
N/ha/yr (Lu and Tian. 2013; Lu et al.. 2012). Tianetal. (2011) applied the Terrestrial
Ecosystem Model and the DLEM model to China using data from 1961-2005. In both
models, China was a C sink during this time period, with the combination of N deposition
and agricultural fertilizer use accounting for 61% of the increase in C sequestration.
Notably, the responsiveness of terrestrial C sequestration to N deposition has declined
since the 1980s as N deposition in China increased, providing evidence that terrestrial
ecosystems are becoming less N limited in China (Tian etal.. 2011). Lu et al. (2012)
conducted a similar analysis, using DLEM to understand how multifactor environmental
change influenced terrestrial C sequestration in China from 1901-2005. Like Tian et al.
(2011). Lu et al. (2012) observed that N deposition increased terrestrial C sequestration in
China throughout much of the late 20th century, but that the responsiveness of terrestrial
C sequestration to N deposition has declined since the 1980s. Moreover, Lu et al. (2012)
reported that all areas of China, aside from some shrublands and portions of western
China, are becoming N saturated.
6-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
6.2.3. Forests
6.2.3.1. Aboveground Processes
The 2008 ISA noted a wide range of forest productivity responses to added N. Responses
to low levels of N were often positive because N availability tends to limit growth in
terrestrial ecosystems (LeBauer and Treseder. 2008). However, forest productivity
responses to higher rates of N addition were neutral or negative [e.g., (Magill et al..
2004)1. The effects of N deposition were variable across species. Conifer species,
particularly at high elevations, were more likely to exhibit negative growth responses or
mortality in response to added N. Conifer species were less likely to demonstrate
increased growth in response to additional N and more often exhibited decreased growth
and increased mortality [e.g., (McNultv et al.. 2005; Beieretal.. 1998; Boxman et al..
1998a)]. Differences between broadleaf and conifer species were especially clear in
long-term N addition experiments: Elvir et al. (2003) observed increased sugar maple
(Acer saccharum) basal area growth in response to long-term (NFL^SC^ (25 kg N/ha/yr
for 10 years) additions, but red spruce (Picea rubens) growth was unchanged. At Harvard
Forest, oak (Quercus velutina, Q. rubra) increased growth in response to chronic N
additions (50 or 150 kg N/ha/yr for 15 years), while red pine (Pinus resinosa) growth
slowed and mortality increased (Magill et al.. 2004). Most empirical observations of the
effects on atmospheric N deposition on forest productivity came from chronic N addition
experiments in temperate forests in the U.S. [e.g., (Pregitzer et al.. 2008; Magill et al..
2004; Elvir et al.. 2003; McNultv et al.. 1996; Aberetal.. 1995)1 and temperate and
boreal forests in Europe (Hvvonen et al.. 2008; Hogberg et al.. 2006; Beier et al.. 1998;
Boxman et al.. 1998a). Empirical analyses of the effects of atmospheric N deposition on
forest productivity in the U.S. were lacking.
Research published since 2008 has reinforced many of the ideas in the 2008 ISA. There is
considerable evidence from deposition gradient studies, forest modeling, and long-term N
addition experiments that N deposition broadly stimulates tree growth and the
productivity of forested ecosystems. Using forest inventory data from the early 1980s
through the mid 1990s for 23 species growing in a region stretching from Ohio to Maine,
Thomas et al. (2010) found that N deposition accelerated growth in 11 species, including
3 of the 4 most abundant species (red maple [Acer rubrum\, sugar maple, and northern
red oak [Quercus rubra]). Negative effects on growth were seen in three species, all of
which were evergreen conifers (red pine [Pinus resinosa], red spruce, northern white
cedar [Thuja occidentalism). All five of the arbuscular mycorrhizal tree species included
in the analysis exhibited increased growth. Xia and Wan (2008) observed positive effects
of added N on growth for both broadleaf and coniferous trees in a meta-analysis, with
June 2018
6-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
broadleaf trees (+73%) more responsive than conifers (+37%). Eight species exhibited
higher mortality rates with increasing N deposition in the Thomas et al. (2010) analysis,
notably several oak species, including northern red oak. Only three species showed
increased survival. Examining forest stand-level responses, Hember et al. (2017)
concluded N deposition had increased forest stand growth in one of the five Canadian
ecozones (Montane Cordillera), decreased it in another (Boreal Plain), and did not
significantly affect stand growth in the remaining three (Pacific Maritime, Boreal Shield,
and Atlantic Maritime). They similarly found differences in forest stand mortality, with N
deposition decreasing mortality in three of the five ecozones (Pacific Maritime, Montane
Cordillera, and Boreal Plain) and no significant effect in the remaining two. In an
analysis of forest inventory data from across the entire eastern U.S. from the 1970s
through early 2000s, Dietze and Moorcroft (2011). as noted in Appendix 5. found N
deposition was linked to decreased tree mortality in 9 of 10 plant functional types and
increased mortality only in the northern midsuccessional hardwoods functional type.
Across a smaller gradient within the Adirondacks, Bedison and McNeil (2009) found a
significant positive overall effect of N deposition on tree growth from 1984-2004, but
positive growth effects only for red maple, balsam fir (Abies balsamea), and red spruce at
the species level. Thus, it appears from these inventory analyses that while tree growth
has generally been enhanced by N deposition over the last several decades, individual
species have exhibited variable responses to N deposition in mortality and growth.
Many of the species exhibiting notable responses in the broad northeastern U.S. inventory
analysis have also been included in long-term simulated N deposition experiments, but
the results in these studies have not always been consistent with the observations of
Thomas et al. (2010). Overstory sugar maple trees increased growth in response to added
N in Michigan lYPregitzer et al.. 2008); N added at 30 kg N/ha/yr] and Maine [(Elviret
al.. 2003); 25 kg N/ha/yr], but mature sugar maple and red maple did not respond in the
Catskills lYLovett et al.. 2013); 50 kg N/ha/yr], Northern red oak increased growth at
Harvard Forest lYFrev et al.. 2014); 50 and 150 kg N/ha/yr] but showed no growth
response at two sites in New York state [(Wallace et al.. 2007); 75 kg N/ha/yr, (Lovett et
al.. 2013); 50 kg N/ha/yr], with increased mortality at one of the New York sites (Wallace
et al.. 2007). A 13-year study in a young forest in West Virginia included the two species
with the most positive growth responses in the Thomas et al. (2010) analysis and found
that N addition (35 kg N/ha/yr of [NH/J2SO4) generally decreased growth of black cherry
(Prunus serotinci) and tulip poplar (Liriodendron tulipifera), although these changes were
not statistically significant (May et al.. 2005). Red pine at Harvard Forest exhibited
decreased growth and higher mortality in response to chronic N additions lYFrev et al..
2014); 50 and 150 kg N/ha/yr], while red spruce showed no growth response in Maine
lYElvir et al.. 2003); 25 kg N/ha/yr], Notably, the rates of N additions in these studies
6-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
were often considerably greater than the typical rates of N deposition observed in the
U.S.
Other studies have also observed similarly mixed results on tree growth and mortality in
U.S. forests. At the same sites in Michigan where long-term N additions increased growth
of mature sugar maple (Prcgitzcr et al.. 2008). N additions decreased the growth and
survival of sugar maple saplings (Talhclm et al.. 2013; Patterson et al.. 2012). Notably,
this negative effect occurred without increase in overstory leaf area that would reduce
light availability or a decrease in soil pH (Talhclm et al.. 2013). More recently, Ibanez et
al. (2016) observed that although small overstory trees (5-10 cm diameter at breast
height) at these sites were growing faster, they were also experiencing increased
mortality. As with sugar maple saplings in Michigan, N deposition appears to be
decreasing growth among northern red oak saplings in the Chicago area, potentially
contributing to the inability of this species to regenerate in that region (Bassirirad et al..
2015). In Vermont, 2 years of N addition at 150 kg N/ha/yr increased growth in four
hardwood species (including sugar maple and northern red oak), decreased growth in one
conifer, and had no effect on two other hardwoods (Finzi. 2009). Allen et al. (2010)
reported higher mortality and a decrease in ectomycorrhizal root tips (Table 6-2). with N
addition (100 kg N/ha/yr) in ectomycorrhizal pinon pine in New Mexico, while no
change in the mortality of arbuscular mycorrhizal juniper. Allison et al. (2010) observed
a 2.5-fold increase in aboveground net primary productivity in response to several years
of N additions (100 kg N/ha/yr) in central Alaska in a recently burned boreal forest. At
two mixed conifer forests in the Sierra Nevada, 2 years of N additions (12 or
24 kg N/ha/yr) had a positive effect on herb community biomass at one site
(24 kg N/ha/yr) and a negative effect at the other site (12 kg N/ha/yr); shrub biomass was
unaffected (Hurteau and North. 2008). Modeling 50 years into the future based on these
results, a lower rate of N deposition (12 kg N/ha/yr) was expected to increase herb
biomass, while the high rate of N deposition (24 kg N/ha/yr) led only to a small increase
in shrub biomass. In combination with a wetter precipitation regime, both N deposition
rates were predicted to increase shrub and herb biomass (Hurteau et al.. 2009). In a
greenhouse study, N additions of up to 120 kg N/ha/yr had no effect on the N fixing tree
black locust (Robinia pseudoacacia) when grown in a monoculture or in competition
with the sawtooth oak (Quercus acutissima), but N additions increased the height and
total biomass of the sawtooth oak when grown in competition with black locust (Luo et
al.. 2014).
6-19
DRAFT: Do Not Cite or Quote

-------
Table 6-2 Growth, productivity, and carbon cycle responses of ectomycorrhizal fungi to nitrogen added via
atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Hoqberq et al. (2010)
Sweden
Scots pine (Pinus
sylvestris)
Addition
100
2
13C labeling of EM
biomarker PLFA
18:2(jo6,9
Decrease
Nasholmetal. (2013)
Sweden
Scots pine (Pinus
sylvestris)
Addition
100
0
(2 weeks)
13C labeling of EM
biomarker PLFA
18:2(jo6,9
Not significant
Parrent and Vilaalvs
(2009)
North Carolina
Loblolly pine (Pinus
taeda)
Addition
112
1
Ectomycorrhizal 18S
RNA expression
Not significant
Hoqberq et al. (2011)
Sweden
Scots pine (Pinus
sylvestris)
Addition
120
20; 15 yr
recovery
Ectomycorrhizal
biomarker 18:2u>6,9
Not significant
Hoqberq et al. (2011)
Sweden
Scots pine (Pinus
sylvestris)
Addition
30, 70
35
Ectomycorrhizal
biomarker 18:2u>6,9
Decrease
Nasholmetal. (2013)
Sweden
Scots pine (Pinus
sylvestris)
Addition
100
0
(2 weeks)
Fine root chitin
concentration
Not significant
Ki0ller et al. (2012)
Denmark
Norway spruce
(Picea abies)
Ambient
27-43
n/a
Mycelium production
Decrease
Bahretal. (2013)
Sweden
Norway spruce
(Picea abies)
Ambient
0.9-24.6
n/a
Mycelium production
Decrease
Bahretal. (2015)
Sweden
Norway spruce
(Picea abies)
Addition
200 (once)
1
Mycelium production
Decrease
Hasselquist et al.
(2012)
Sweden
Scots pine (Pinus
sylvestris)
Addition
20, 100
6
Mycorrhizal respiration
Low dose: increase
Hiqh dose: decrease
June 2018
6-20
DRAFT: Do Not Cite or Quote

-------
Table 6 2 (Continued): Growth, productivity, and carbon cycle responses of ectomycorrhizal fungi to nitrogen
added via atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Vallack et al. (2012)
Sweden
Scots pine (Pinus
sylvestris)
Addition
100
2
Mycorrhizal respiration
Decrease
Allen et al. (2010)
New Mexico
Pinon pine (Pinus
edulis) and
one-seeded juniper
(Juniperus
monospermum)
Addition
100
7+
Mycorrhizal root tip
colonization (by
ectomyorrhizal and
arbuscular
mycorrhizae)
Decrease in
ectomycorrhizal root
tips; no change in
arbuscular
mycorrhizal root tips
Pritchard et al. (2014)
North Carolina
Loblolly pine (Pinus
taeda)
Addition
112
6
Mycorrhizal root tip
production
Decrease
Pritchard et al. (2014)
North Carolina
Loblolly pine (Pinus
taeda)
Addition
112
6
Mycorrhizal root tip
survival (deep soil)
Increase
Kou et al. (2017)
China (sub-
tropical)
Slash pine (Pinus
eiiiottii)
Addition
40, 120
2
Mycorrhizal
(ectomycorrhizal)
survival
Increase at both low
and high additions,
except for deeper soil
and dichotomous
mycorrhizae
Parrent and Vilqalvs
(2009)
North Carolina
Loblolly pine (Pinus
taeda)
Addition
112
1
Root ammonium
transport gene
expression
Not significant
Garcia et al. (2008)
North Carolina
Loblolly pine (Pinus
taeda)
Addition
112
2
Root colonization (%)
Increase
Diaz et al. (2010)
Spain
Aleppo pine (Pinus
haiepensis)
Addition
35, 60, 120 mg/
plant
1
Root colonization (%)
Decrease
Ki0ller et al. (2012)
Denmark
Norway spruce
(Picea abies)
Ambient
27-43
n/a
Root colonization (%)
Decrease
Kou et al. (2015)
China
Slash pine (Pinus
eiiiottii)
Addition
40, 120
2
Root colonization (%)
Low dose: not
significant
Hiah dose: increase
June 2018
6-21
DRAFT: Do Not Cite or Quote

-------
Table 6 2 (Continued): Growth, productivity, and carbon cycle responses of ectomycorrhizal fungi to nitrogen
added via atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Parrent and Vilqalvs
(2009)
North Carolina
Loblolly pine (Pinus
taeda)
Addition
112
1
Root monosaccharide
transport gene
expression
Not significant
Allison et al. (2008)
Alaska
Black spruce (Picea
mariana)
Addition
140
5
Sporocarp abundance
Decrease
Allen et al. (2010)
New Mexico
Pinon pine (Pinus
edulis) and one-
seeded juniper
(Juniperus
monospermum)
Addition
100
7+
Sporocarp abundance
Decrease
Gillet et al. (2010)
Switzerland
Norway spruce
(Picea abies)
Addition
150
12
Sporocarp abundance
Decrease
Hasselquist et al.
(2012)
Sweden
Scots pine (Pinus
sylvestris)
Addition
20, 100
6
Sporocarp abundance
Low dose: not
significant
Hiqh dose: decrease
Hasselauist and
Hoqberq (2014)
Sweden
Scots pine (Pinus
sylvestris)
Addition
110
20, 15 yr
recovery
Sporocarp abundance
Not significant
Hasselquist and
Hoqberq (2014)
Sweden
Scots pine (Pinus
sylvestris)
Addition
20, 100
6
Sporocarp abundance
Low dose: not
significant
Hiqh dose: decrease
Hasselquist and
Hoqberq (2014)
Sweden
Scots pine (Pinus
sylvestris)
Addition
35, 70
40, 2 yr
recovery for
70-kg
treatment
Sporocarp abundance
Decrease
EM = ectomycorrhizal fungi; ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; PLFA = phospholipid fatty acids; RNA = ribonucleic acid; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
A number of studies on the effects on N deposition or chronic N additions have been
conducted in temperate and boreal forests in Europe, and these studies have more
consistently found positive effects of N on tree growth (e.g., (Nair et al.. 2016; Ferretti et
al.. 2014; Solberg et al.. 2009). Ferretti et al. (2014) observed a positive relationship
between N deposition and tree growth across 26 forest plots in Italy. A positive effect of
N deposition on tree growth was also found on 135 Swiss long-term observation plots for
mature beech and spruce from the mid 1980s through late 2000s (Braun et al.. 2010).
Further N additions decreased tree growth at three of seven experimental sites, yet these
changes were linked to P deficiencies (Braun et al.. 2010). Eastaugh et al. (2011) took a
more complex approach to understand growth trends of Norway spruce (Picea abies) in
the Austrian National Forest Inventory from 1960-2008. Using the Biome-BGC model to
isolate growth trends from variation in climate, the authors found a positive relationship
between tree growth and N deposition. Using data from long-term N addition
experiments in Norway spruce and Scots pine (Pinus sylvestris) forests throughout
Sweden and Finland, Hwonen et al. (2008) found increased tree growth at 11 of 12 sites.
The effects of added N on forest productivity appear to persist for decades and carry over
through major ecosystem disturbances. From et al. (2015) studied young Norway spruce
and Scots pine forests planted in the late 1990s after the original forests were clear-cut.
The original forests had received N additions of 150 kg N/ha/yr (as NH4NO3) either twice
(in 1977 and 1985), once (1985), or never. Tree height growth of the young forests was
positively related to the amount of previous N additions, with significantly greater height
growth in the forests that had received two N additions. Foliar N concentrations were also
significantly higher in the forests that had received N additions (From et al.. 2015).
Biomass of understory shrubs and bryophytes decreased in response to N additions
(100 kg N/ha/yr for 6 years) on boreal forest islands in Sweden, although one of three
shrub species increased in growth (Wardle et al.. 2016).
The 2008 ISA tied greater tree growth and forest productivity to increases in foliar N,
photosynthesis, and gross primary productivity. In addition, foliar N was identified as a
sensitive indicator of changes in forest N availability (Aberetal.. 1989). Since the 2008
ISA, increases in foliar N continue to have been linked to N deposition in forests in the
U.S., Europe, and Asia in pollution gradient studies (Sardans et al.. 2016; Talhelm et al..
2012; Jones et al.. 2011; Cox et al.. 2010; Thimonier et al.. 2010; Fenn et al.. 2008) and N
addition experiments (Du. 2017; Fusaro et al.. 2017; Gilliam et al.. 2016a; Wardle et al..
2016; Fowler et al.. 2015; Du and Fang. 2014; Lovett et al.. 2013; Talhelm et al.. 2013;
Lovett and Goodale. 2011; Talhelm etal.. 2011; Allen et al.. 2010; Braun et al.. 2010).
For instance, Cox et al. (2010) observed a 40% increase in foliar N (from 15 mg/g to
21 mg/g) in Scots pine forests in Germany and the U.K. along an N deposition gradient of
4.6 to 28 kg N/ha/yr. In addition to trees, forest bryophytes and lichens also show
increased tissue N concentrations when exposed to additional N (Maaroufi et al.. 2016;
6-23
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
McMurrav et al.. 2015; McMurrav et al.. 2013; Gundale et al.. 2011; Fenn et al.. 2008;
Glavich and Geiser. 2008). Increased foliar N is not universally observed in N addition
studies [e.g., (Zhang et al.. 2017; Zhang et al.. 2015c; Lovett et al.. 2013)1 or along N
deposition gradients (Watmough and Meadows. 2014). but a meta-analysis found that N
additions generally increase foliar N in trees (Lu etal.. 2011b).
Although there are clear links between N deposition and increased foliar N and between
higher foliar N and increased photosynthesis, there is only limited evidence that chronic
N deposition directly increases leaf-level photosynthesis in forests. For instance, in
long-term simulated N deposition experiments in Massachusetts and Maine, only one of
four species exhibited increased photosynthesis with N additions (Elvir etal.. 2006;
Bauer et al.. 2004). Research on photosynthesis since the 2008 ISA has been similarly
mixed. Talhelm et al. (2011) did not observe a significant increase in either
photosynthesis or canopy leaf area in mature sugar maple trees at four sites in Michigan
in response to N additions (30 kg N/ha/yr as NaNC>3 for 14 years). Using a
canopy-applied N treatment at two Swiss forests, Wortman et al. (2012) found that N
additions improved photosynthetic processes in oak at one site, but did not significantly
influence photosynthetic parameters in spruce or beech at the other site. In Swedish
boreal forests, higher tissue N concentrations resulting from N additions (20 or
100 kg N/ha/yr for 4 years) increased the amount of 13CC>2 assimilated into shrub
(Vaccinium) foliage but decreased the amount of C assimilated by bryophytes
(Hasselquist et al.. 2016). At a broader scale, Fleischer et al. (2013) analyzed the effect of
N deposition on photosynthesis using observations from 80 forest eddy covariance C flux
measurement sites predominantly located in the eastern U.S. and western Europe. In this
data set, canopy photosynthetic capacity was positively correlated to N deposition for
conifer forests when N deposition was below an observed critical load of approximately
8 kg N/ha/yr (Fleischer et al.. 2013). Most of the stimulus provided by N deposition
occurred in boreal forests, with little to no influence of N deposition in broadleaf forests
or forests in the temperate climate zones (Fleischer et al.. 2013). Leonardi et al. (2012)
compiled a global data set of published tree ring 13C chronologies and found that the rate
of N deposition had a significant linear effect on intrinsic water-use efficiency (iWUE)
and 13C discrimination (A13C) in conifers and a quadratic effect in angiosperms. This
change in 13C discrimination suggests that the additional N either stimulated
photosynthesis or stimulated growth to the extent that the trees became more water
limited.
The apparent contradiction between the wide observations that N deposition can increase
both tree growth and foliar N concentrations and the inconsistent effects of N deposition
on leaf-level photosynthesis could be explained by two other physiological responses to
N. First, plants can store excess N in foliage as free amino acids and other
6-24
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
physiologically inactive forms of N (Bauer etal.. 2004). Increases in foliar amino acid
concentrations, particularly arginine, have been observed in forests in chronic N addition
experiments (Bauer et al.. 2004) and along N deposition gradients (Braun et al.. 2010). In
these cases, the added foliar N would have no impact on photosynthesis. Second, plants
can respond to added N by altering C allocation. As noted in the 2008 ISA, higher N
availability often causes trees to allocate less photosynthate to roots and belowground
processes and more toward aboveground growth (Vicca et al.. 2012; Hogbcrg et al..
2010; Janssens et al.. 2010; Litton et al.. 2007; Aerts and Chapin. 1999; Minnich et al .
1995). Research since the 2008 ISA has provided further evidence that N deposition is
likely capable of causing such a shift in forests. In a greenhouse experiment, N additions
(2, 5, and 10 kg N/ha/yr) resulted in more biomass allocated to leaves than to roots at all
four N addition levels for oak and black locust trees. At the same Michigan sites where
Talhelm etal. (2011) observed that chronic N additions had no effect on leaf-level
photosynthesis or stand leaf area, the combination of increased aboveground growth, no
change in root production or respiration, and decreased mycorrhizal abundance led the
authors to conclude that a change in C allocation away from mycorrhizae was the likely
cause for enhanced tree growth. More broadly, Vicca et al. (2012) found that across
49 forest sites, higher site fertility increased both the fraction of biomass that was
allocated aboveground and the fraction of photosynthate that was allocated toward
biomass production. The increase in C allocated to biomass production was thought to
most likely result from less C being allocated to root symbionts such as mycorrhizae
(Vicca et al.. 2012). Likewise, Janssens et al. (2010) conducted a meta-analysis of
20 forest N addition experiments and found that N additions decreased the fraction of C
allocated to belowground processes but did not affect root biomass. As an example of
this, Hogberg et al. (2010) conducted a short-term (~2 hour) 13CC>2 canopy labeling
experiment in a young Scots pine forest in northern Sweden to determine how N
additions (two 100 kg N/ha additions over 2 years) altered the allocation of recent
photosynthate belowground. In the 3 weeks following the canopy labeling, the total soil
respiratory efflux of the 13C label was 62% lower in the plots that had received the N
additions. Other observed physiological responses to N include delayed bud burst (De
Barba etal.. 2016). tree ring width, and xylem conduit density (Borghetti et al.. 2017).
The effects of N addition on forest productivity have the potential to be moderated by
increases in herbivory, particularly among insects (Throop and Lerdau. 2004). For
instance, Andersen et al. (2010) observed that N additions increased foliar N and
photosynthetic rates within a tropical forest in Panama, but that tree growth rates were
unaffected because of increased herbivory. At Harvard Forest in Massachusetts, northern
red oak regeneration declined with N additions (50 and 100 kg N/ha/yr) (Bogdziewicz et
al.. 2017). Acorn production increased, but so did weevil infestation of the acorns, and
germination rates declined. In a Pennsylvania common garden experiment with northern
6-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
red oak, saplings receiving N additions (200 kg N/ha/yr) tended to have greater herbivore
damage, but the effects of N addition on herbivory and the link between herbivory and
decreased growth both varied by the type of herbivore, the tree lineage (breeding family),
and the study location (Cha et al.. 2010). Similarly, Jones et al. (2008) and Jones et al.
(2011) observed that the most abundant herbivore of bracken fern (Pteridium aquilinum)
and a beetle herbivore of California black oak (Quercus kelloggii) were increased by N
addition at a heavily polluted site, but not at a drier, less polluted site. However, there
was a positive correlation between leaf NOs concentrations and the abundance of several
groups of black oak herbivores.
6.2.3.2. Belowground Processes
The 2008 ISA analysis of how belowground processes in forests reacted to N deposition
focused on fine root dynamics, aboveground litter inputs, decomposition, soil respiration,
and soil C. Johnson (2001) had found a significant increase in forest soil C in response to
N additions as part of a meta-analysis, but only a single long-term N addition field study
had observed a significant increase in forest soil C KPrcgitzcr et al.. 2008); see
Appendix 41. For fine roots, Nadelhoffer (2000) hypothesized that N deposition would
decrease biomass, but stimulate turnover and production. The 2008 ISA found little
evidence with which to evaluate this hypothesis. There was more available research on N
effects on decomposition and soil respiration. As noted earlier, Knorretal. (2005)
observed in a meta-analysis that N additions increased decomposition at sites receiving
low rates of ambient N deposition (<5 kg N/ha/yr), but N additions suppressed
decomposition at sites receiving moderate rates of N deposition (5-10 kg N/ha/yr).
Research completed since the 2008 ISA has advanced our understanding of how N
affects belowground processes in forests. In two meta-analyses, Liu and Greaver (2010)
did not find a consistent effect of N additions on forest fine root production, while Li et
al. (2015) observed a decrease (-13.5%) in forest fine root biomass. Broadly, some of the
results discussed in the 2008 ISA that appeared inconsistent at the time were likely a
reflection of how N additions interacted with other ecological processes. In young,
rapidly expanding forests, N addition may increase root biomass as a consequence of an
overall enhancement of plant growth (Janssens et al.. 2010). For instance, N additions
(150 kg N/ha/yr for 3 years) to a young bamboo forest increased fine root growth by
>30%, even though this forest already received high rates of atmospheric N deposition
[95 kg N/ha/yr of wet deposition; (Tu et al.. 2015)1. However, as noted in the
"Aboveground Processes" section (Appendix 6.2.3.1). increases in N deposition tend to
decrease the proportion of C allocated to roots relative to aboveground growth (Li et al..
2015; Vicca et al.. 2012; Janssens et al.. 2010; Litton et al.. 2007; Minnich etal.. 1995).
June 2018	6-26	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Thus, in more mature forests, root biomass and production are not consistently affected
by N deposition. For instance, there were no effects of chronic N additions on fine root
biomass in four mature hardwood forests in Michigan [30 kg N/ha/yr for over 15 years;
(Burton et al.. 2012)1. and in two mature mixed oak (Quereus) stands at Harvard Forest
[50 or 150 kg N/ha/yr for over 20 years (Frcv et al.. 2014)1. Likewise, N addition
(50 kg N/ha/yr) did not cause significant changes in root biomass in a tropical forest in
Puerto Rico (Cusack etal.. 2010). Root turnover did not respond to N addition
(30 kg N/ha/yr for 3 years) across 13 successional (20-40 years old) and mature
(>90 years old) hardwood stands in central New Hampshire (Kang et al.. 2016). In
contrast with the mixed oak stands, fine root biomass decreased in two red pine stands at
Harvard Forest [50 or 150 kg N/ha/yr for over 10 years; (Frev etal.. 2014)1. and root
biomass also decreased with N additions (50 or 150 kg N/ha/yr for 3 years) in a
subtropical, broadleaf forest in China (Peng et al.. 2017).
Observations of how other belowground processes respond to N additions provide further
evidence of an increase in the ratio of tree C allocated to aboveground growth and
productivity versus the C allocated belowground. In a meta-analysis of forest soil
respiration responses to N addition, Janssens et al. (2010) noted that N additions
decreased soil respiration overall, with a portion of this effect caused by a decrease in
autotrophic respiration. Within forests, autotrophic respiration may make up more than
50% of total soil respiration (Hasselquist et al.. 2012). Much of this autotrophic
respiration is mycorrhizal respiration, with the C allocated to and respired by
mycorrhizae estimated to account for 9 to 34% of forest soil respiration (Hasselquist et
al.. 2012; Heinemever et al.. 2007).
Mycorrhizal fungi have long been observed to be sensitive to increased forest N
availability, particularly ectomycorrhizae because these fungi have direct roles in plant N
acquisition, and their production of aboveground fruiting bodies (sporocarps) makes it
easier to observe changes in abundance (Treseder. 2004; Lilleskov et al.. 2002; Wallenda
andKottke. 1998). As noted in Appendix 4. when N supply increases, C allocation to
ectomycorrhizal fungi decreases, and their abundance and activity decline. In a
meta-analysis, Li et al. (2015) found that N additions decreased mycorrhizal colonization
of fine root tips by 19% (n = 12). In contrast to the widespread increases in aboveground
tree growth, studies of ectomycorrhizal growth and productivity responses to added N
have documented nearly universally negative or neutral effects on metrics such as
mycorrhizal root colonization, sporocarp abundance, and the abundance of the fungal
lipid biomarker 18:2co6,9 in the soil (Table 6-2). Although this research is consistent,
these studies have limitations because nearly all of this research has been conducted on
conifer species and most of the negative effects occur in studies using unrealistically high
rates of N addition (>100 kg N/ha/yr). However, two studies of ectomycorrhizae in
6-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Norway spruce forests in Europe have documented significant declines along gradients of
ambient N deposition (Bahr etal.. 2013; Kioller et al.. 2012). In particular, (Ki oiler et al..
2012) observed an 80 and 90% decrease in ectomycorrhizal root tip abundance and
mycelial production, respectively, across a canopy throughfall N deposition gradient of
27 to 43 kg N/ha/yr in Denmark. Morrison et al. (2016) also observed a decline in the
relative abundance of ectomycorrhizal fungal DNA within the soil of the N addition (50
or 150 kg N/ha/yr for 25 years) plots at the oak (6>?it'/x'?w)-dominatcd Harvard Forest in
Massachusetts.
Given these decreases in ectomycorrhizal growth and productivity in response to added
N, it is not surprising that the amount of C allocated belowground by plants to
mycorrhizae can be sensitive to N availability. Hogberg (2012) twice added N
(100 kg N/ha) over a period of two growing seasons to boreal Scots pine forests in
Sweden as part of a short (~2 hour) 13C canopy labeling experiment. In the 5 weeks
following the labeling, 48% less of the 13C label was found in an ectomycorrhizal fungal
lipid biomarker in the soil of the N amended plots. Hasselquist et al. (2012) found that in
Scots pine forests in northern Sweden, 6 years of N additions at a rate of 100 kg N/ha/yr
decreased respiration from ectomycorrhizae by 40%, but N additions at a rate of
20 kg N/ha/yr increased this respiration by 120%. However, the low N addition did not
change the fractional contribution of ectomycorrhizae to total soil respiration. The low N
addition treatment also had no effect on ectomycorrhizal sporocarp production, but the
high N addition treatment nearly eliminated sporocarp production [99% decrease;
(Hasselquist et al.. 2012)1. The availability of N also appears to influence the transfer of
N from mycorrhizal fungi to plants. In a boreal Scots pine forest in Sweden, Nasholm et
al. (2013) observed that a single N dose (100 kg N/ha) shifted the dominant sink for a 15N
tracer added 2 weeks later from the cytoplasm of ectomycorrhizal fungi and other soil
microorganisms to the pine foliage.
Although ectomycorrhizae are important in many high latitude and temperate
ecosystems, particularly forests, most species of terrestrial plants form arbuscular
mycorrhizae (Rillig. 2004). The impact of N deposition on arbuscular mycorrhizae has
received less research attention, perhaps partly because these fungi are best known for
their role in plant P acquisition (Rillig. 2004). Arbuscular mycorrhizae community
composition and production can be sensitive to added N IYvan Diepen et al.. 2010;
Egerton-Warburton and Allen. 2000); Table 6-31. but these effects may not be consistent.
van Diepen et al. (2010) reviewed eight previous studies of how intraradical (within root)
and extraradical arbuscular mycorrhizal biomass responded to N additions,
predominantly in forests, and found inconsistent effects. Intraradical biomass
significantly declined in response to N in three studies and increased in two studies,
including in the work of (Garcia et al.. 2008). In comparison, extraradical biomass was
6-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
either unresponsive or declined (van Diepen et al.. 2010). Aside from the work of Garcia
et al. (2008). other researchers since 2008 have found either neutral or negative effects on
metrics of arbuscular mycorrhizal biomass and abundance (Table 6-3). Observations of
decreased arbuscular mycorrhizal biomass span from a subtropical broadleaf evergreen
forest in China that experienced a single year of N additions (25 or 50 kg N/ha/yr) (Shi et
al.. 2016a) to a subalpine Englemann spruce (Picea engelmannii) forest in Rocky
Mountain National Park that had received more than 15 years of N additions
(25 kg N/ha/yr) (Boot et al.. 2016).
Notably, the arbuscular mycorrhizal species studied by Garcia et al. (2008) were all
understory or subdominant canopy species growing beneath ectomycorrhizal loblolly
pines. The increased growth of these larger trees in response to added N may have
imposed other resource limitations on the arbuscular mycorrhizal plant species. In
general, the response of arbuscular mycorrhizae to N additions may depend upon the
relative availability of P or other nutrients. Johnson et al. (2003) found N additions
decreased arbuscular mycorrhizae when the soil N:P ratio was low (i.e., P-rich soils),
while increasing arbuscular mycorrhizae under high soil N:P (i.e., P-poor soils). Where P
is more limiting, plants may allocate more carbon to arbuscular mycorrhizae in order to
acquire it (Egerton-Warburton et al.. 2007; Johnson et al.. 2003). Thus, the varying
availability of soil P may explain in part the seemingly inconsistent response of
arbuscular mycorrhizae to N additions.
6-29
DRAFT: Do Not Cite or Quote

-------
Table 6-3 Growth, productivity, and carbon cycle responses of arbuscular
mycorrhizal fungi to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or
Addition
Nitrogen
Addition
Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
van Dieoen et al.
(2010)
Michigan
(four
sites)
Sugar maple
(Acer
saccharum)
Addition
30
13
Extraradical
biomass
production
Decrease
van Diepen et al.
(2010)
Michigan
(four
sites)
Sugar maple
(Acer
saccharum)
Addition
30
13
Extraradical
biomass
(16:1 oj5c
abundance)
Decrease
Chen et al. (2014)
China
Steppe
grassland
Addition
100
6
Hyphal length
Decrease
van Dieoen et al.
(2010)
Michigan
(four
sites)
Sugar maple
(Acer
saccharum)
Addition
30
13
Intraradical
biomass
(16:1 oj5c
abundance)
Decrease
Garcia et al. (2008)
North
Carolina
Loblolly pine
(Pinus
taeda)
Addition
100
1-2
Root
colonization
(%)
Increase
Mandvam and
Jumpponen (2008)
Kansas
C3 and C4
grasses
Addition
100
3
Root
colonization
(%)
Not
significant
Van Der Heiiden et
al. (2008)
Holland
Dune
grasses
Addition
100
1
Root
colonization
(%)
Decrease
Camenzind et al.
(2014)
Ecuador
Evergreen
tropical
forest
Addition
50
3
Root
colonization
(%)
Decrease
Chen et al. (2014)
China
Steppe
grassland
Addition
100
6
Root
colonization
(%)
Not
significant
Garcia et al. (2008)
North
Carolina
Loblolly pine
(Pinus
taeda)
Addition
100
1-2
Soil glomalin
content
Not
significant
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only
statistically significant effects are listed as increases or decreases.
1	Aside from mycorrhizal respiration, the other dominant component of autotrophic soil
2	respiration is root respiration. Although Janssens et al. (2010) observed a decrease in
June 2018
6-30
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
autotrophic soil respiration in the presence of N, this decrease was not apparently related
to changes in root respiration. In a long-term N addition experiment (30 kg N/ha/yr for
+15 years) in four northern hardwoods forests, Burton et al. (2012) found that although
root N concentration was a strong predictor of the root respiration rate and N additions
had increased root N concentrations, N additions had not increased root respiration
because the added N changed the relationship between root N and respiration. Drake et
al. (2008) made a similar observation in loblolly pine forests in North Carolina: N
additions (100 kg N/ha/yr for 2 years) increased fine root N concentrations but did not
impact fine root respiration rates, resulting in an altered tissue N respiration relationship.
Frev et al. (2014) also reported that root respiration was unresponsive to long-term N
additions (50 or 150 kg N/ha/yr) in oak forests and pine forests in Massachusetts.
However, Hasselquist et al. (2012) found that a relatively modest N addition rate
(20 kg N/ha/yr for 6 years) increased both root and mycorrhizal respiration in a boreal
forest in Sweden that received relatively low rates of ambient N deposition
(<5 kg N/ha/yr). Thus, it appears that C allocation to mycorrhizae is more sensitive to N
availability than C allocation to root respiration.
In a meta-analysis of the effects of N additions on microbial biomass, (Treseder. 2008)
found that changes in soil respiration were significantly and positively correlated with the
response of microbial biomass. Mycorrhizal fungi are major components of forest soil
microbial communities, with Hogberg et al. (2010) estimating that ectomycorrhizal
mycelium made up 39% of total soil microbial biomass in a Swedish boreal forest. At
Harvard Forest, (Morrison et al.. 2016) observed that 59-72% of all fungal operational
taxonomic units (OTUs) belonged to ectomycorrhizal fungi. Given the widespread
negative effects of added N on mycorrhizal fungi, the results of Treseder (2008)
meta-analysis and the broadly negative or neutral effects of N additions on microbial
biomass in the studies published since 2008 are unsurprising (Table 6-4).
6-31
DRAFT: Do Not Cite or Quote

-------
Table 6-4 Abundance and carbon cycle responses of forest soil
microorganisms and soil invertebrates to nitrogen added in
experimental treatments.
Ambient
Deposition
Study	or
Reference	Location Vegetation Addition
Nitrogen	Effect of
Addition Rate	Additional
(kg N/ha/yr) Duration (yr) Endpoint Nitrogen
Treseder (2008)	Meta-analysis
Mostly Addition	1 -600
boreal and
temperate
forests
0.5-57 Bacterial Not
biomass significant
Zhao et al. (2014a) China (Tibetan Spruce-fir Addition	250	4	Bacterial Decrease
plateau)	(Picea	biomass
asperata,
Abies
faxoniana)
Hesse et al. (2015)
Michigan (Ml
gradient)
Northern
hardwood
forests (Acer
saccharum)
Addition
30
16
Bacterial
biomass
Not
significant
Treseder (2008)
Meta-analysis Mostly
boreal and
temperate
forests
Addition
1-600
0.5-57
Fungal
biomass
Not
significant
Zhao et al. (2014a) China (Tibetan Spruce-fir Addition
plateau)	(Picea
asperata,
Abies
faxoniana)
250
Fungal	Decrease
biomass
Hesse et al. (2015) Michigan (Ml Northern Addition
gradient)	hardwood
forests (Acer
saccharum)
30
16	Fungal	Not
biomass significant
Enowashu et al.
(2009)
Germany Norway Subtraction
spruce
(Picea
abies)
9.7 (-21) 16 (recovery) Fungal
biomass
(ergosterol)
Increase
Bebber et al. (2011) U.K.
Broadleaf Addition
temperate
forest
(Fraxinus-
Acer, Fagus)
2.8
1	Fungal	Not
mycelium significant
growth
van Diepen et al.
(2017)
Massachusetts Temperate, Addition
(Harvard
Forest)
mixed
hardwood
forest
(Quercus
veiutina,
Quercus
rubra)
50, 150
28
Fungal
mycelium
growth
Not
significant
overall;
varied by
fungal
isolate
June 2018
6-32
DRAFT: Do Not Cite or Quote

-------
Table 6-4 (Continued): Abundance and carbon cycle responses of forest soil
microorganisms and invertebrates to nitrogen added in
experimental treatments.



Ambient







Deposition
Nitrogen


Effect of

Study

or
Addition Rate


Additional
Reference
Location
Vegetation
Addition
(kg N/ha/yr)
Duration (yr)
Endpoint
Nitrogen
Allison et al. (2008)
Alaska
Boreal forest
Addition
140
5
Fungal
Decrease


(Picea



sporocarp



mariana)



biomass

Linetal. (2017)
China
Subtropical
Addition
47
10 mo
Invertebrate
Increase


deciduous



density (pre-
on


and



dominantly
coniferous


coniferous



Collembola
litter; not


forests



and Acari) on
significant






decomposing
on






litter
deciduous







litter
Allison et al. (2008)
Alaska
Boreal forest
Addition
140
5
Microbial
Not


(Picea



biomass
significant


mariana)





Treseder (2008)
Meta-analysis
Mostly
Addition
1-600
0.5-57
Microbial
Decrease


boreal and



biomass



temperate







forests





Keeler et al. (2009)
Minnesota
Temperate
Addition
100
5
Microbial
Not

(Cedar Creek)
forests



biomass
significant


(Quercus







ellipsoidalis,







Pinus







strobus) and







Grassland





van Dieoen et al.
Michigan (Ml
Northern
Addition
30
12
Microbial
Decrease
(2010)
gradient)
hardwood



biomass



forests (Acer







saccharum)





Hobbie et al. (2012)
Minnesota
Oak and
Addition
100
5
Microbial
Not

(Cedar Creek)
pine forests



biomass
significant


(Quercus







ellipsoidalis,







Pinus







strobus)





Zhao et al. (2014a)
China (Tibetan
Spruce-fir
Addition
250
4
Microbial
Decrease

plateau)
(Picea



biomass



asperata,







Abies







faxoniana)





Allison et al. (2010)
Alaska
Boreal forest
Addition
114
7
Microbial
Decrease


(Picea



biomass C



mariana)





Wana et al. (2015a)
China
Subtropical
Addition
50, 100
8
Microbial
Low dose:

(southern)
pine forest



biomass C
increase


(Pinus




Hiah dose:


massoniana)




not







significant
June 2018
6-33
DRAFT: Do Not Cite or Quote

-------
Table 6-4 (Continued): Abundance and carbon cycle responses of forest soil
microorganisms and invertebrates to nitrogen added in
experimental treatments.



Ambient






Deposition
Nitrogen

Effect of

Study

or
Addition Rate

Additional
Reference
Location
Vegetation
Addition
(kg N/ha/yr) Duration (yr)
Endpoint
Nitrogen
Wana et al. (2015a)
China
Subtropical
Addition
50, 100 8
Microbial
Not

(southern)
pine-


biomass C
significant


broadleaf






forest (Pinus






massoniana)




Wanq et al. (2015a)
China
Subtropical
Addition
50, 100, 150 8
Microbial
Low, mid

(southern)
broadleaf


biomass C
dose: not


forests



significant






Hiah dose:






decrease
Pena et al. (2017)
China
Mid-
Addition
50, 150 2+
Microbial
Decrease

(Sichuan
subtropical,


biomass C


Province)
evergreen,






broadleaf






forest




Lin et al. (2017)
China
Subtropical
Addition
47 10 mo
Microbial
Not


deciduous


biomass on
significant


and


decomposing



coniferous


litter



forests




Kana etal. (2016)
New
Northern
Addition
30 2,3
Microbial
Not

Hampshire
hardwood


respiration
significant


forests






(Acer






saccharum,






Fagus






grandifolia)




Gillet et al. (2010)
Switzerland
Norway
Addition
150 12
Saprobic
Increase


spruce


fungal
and


(Picea


sporocarp
decrease


abies)


abundance
(N x yr)
van Dieoen et al.
Michigan (Ml
Northern
Addition
30 12
Saprotrophic
Not
(2010)
gradient)
hardwood


fungal
significant


forests (Acer


biomass



saccharum)




C = carbon; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only
statistically significant effects are listed as increases or decreases.
1	However, forest soil microbial communities are taxonomically and functionally diverse
2	(see Appendix 6.3.3.3) and have exhibited varying responses to added N (Zhang et al..
3	2015c; Treseder. 2008). For instance, Wang et al. (2015a) added N (50, 100, or
4	150 kg N/ha/yr for 9 years) to pine, broadleaf, and mixed pine-broadleaf forests in
5	southern China; the 50 kg N treatment increased microbial biomass C in the pine forests
June 2018	6-34	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
and the 150 kg N treatment decreased microbial biomass C in the broadleaf forests, but
there were no other significant treatment effects. Elsewhere in China, a single year of N
additions (25 or 50 kg N/ha/yr) in a subtropical broadleaf evergreen and temperate
broadleaf deciduous forest had no effect on total microbial biomass in either forest, but
decreased the abundance of Gram-negative bacteria, actinomycetes, and saprotrophic
fungi in the subtropical evergreen forest (Shi et al.. 2016a'). The addition of
100 kg N/ha/yr to a regenerating Alaskan boreal forest decreased the microbial biomass
and the C:N ratio of microbial biomass (Allison et al.. 2010; Allison et al.. 2008). while
N additions (50 kg N/ha/yr) in Indiana mixed hardwood forests decreased microbial
biomass without affecting the microbial C:N ratio (Midglev and Phillips. 2016). In
Sweden, the addition of 100 kg N/ha/yr to boreal forests for 6 years decreased bacterial
biomass, but did not significantly affect fungal biomass (Wardle et al.. 2016). The
opposite result was observed in pine and hardwood stands at Harvard Forest and in
subalpine spruce-fir forests in Rocky Mountain National Park (RMNP). There, N
additions decreased fungal biomass, but bacterial biomass was unaffected (50 and 150 kg
N/ha/yr for 14 years at Harvard Forest; 25 kg N/ha/yr for 17 years in RMNP) (Boot et al..
2016). In Michigan, long-term (+15 years) N additions (30 kg N/ha/yr) to northern
hardwood forests decreased the abundance of actinobacteria in the surface mineral soil,
but not in the soil organic horizon (Eisenlord and Zak. 2010). Relative to mycorrhizal
fungi, there is less information available regarding the growth and biomass responses of
individual domains and other taxonomic groupings within microbial communities
Table 6-4). Of this finer scale information, most observations are for fungi (sometimes
including mycorrhizal fungi), with positive, neutral, and negative effects of added N.
Atmospheric N deposition may affect forest microbial communities and litter
decomposition by altering litter chemistry through shifts in forest community
composition, increasing litter N concentrations, or changing the secondary chemistry of
litter. The two dominant forms of forest plant litter, leaves and fine roots, can each be
altered by N additions (Xia etal.. 2015). As noted in the aboveground processes section
of this appendix, there is abundant evidence that N deposition can increase foliar N
concentrations in trees and there is also a large body of research that the increase in foliar
N concentrations will result in increased leaf litter N concentrations (Aerts. 1996). Since
the 2008 ISA, there have been new observations of increased leaf litter N concentrations
both in long-term N addition experiments (van Diepen et al.. 2015; Xia et al.. 2015; Zak
et al.. 2008) and along an N deposition gradient [6.8-11.8 kg N/ha/yr; (Talhelm et al..
2012)1. Increases in leaf litter N concentrations with greater N deposition are not
universal [e.g., (Watmough and Meadows. 2014)1. but in a meta-analysis, van Diepen et
al. (2015) observed a >20% increase in tree leaf litter N concentrations in response to N
additions. Other changes in forest litter chemistry likely involve changes in more
complex physiological and biogeochemical mechanisms (Du and Fang. 2014). In an old
6-35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
growth boreal forest in China, N additions (20, 50, or 100 kg N/ha/yr for 3 years)
decreased leaf litter P concentrations, an effect that could have been caused by changes in
either physiological or biogeochemical processes (Du and Fang. 2014). In a
meta-analysis, van Diepen et al. (2015) found that simulated N deposition significantly
decreased leaf litter concentrations of P, calcium, manganese, aluminum, and zinc, but
did not find significant changes in concentrations of potassium (K), magnesium, boron,
iron, or copper. It is not entirely understood how all of these changes in elemental
composition impact the composition and function of soil microbial communities, but
manganese and calcium are needed for the production of extracellular enzymes that
degrade lignin (van Diepen et al.. 2015).
There is a large volume of research about how major biochemical constituents of plant
litter such as lignin, cellulose, condensed tannins, and phenolics impact leaf litter
decomposition, but less information about how N additions change the abundance of
these compounds. Xia et al. (2015) researched the effects of long-term N additions
(30 kg N/ha/yr for +15 years) on sugar maple leaf litter and fine root chemistry in four
mature northern hardwood forests in Michigan. The N additions increased concentrations
of condensed tannins in leaf litter and in fine roots. At three of the four sites, N additions
increased the fraction of nonstructural cell wall material and decreased both cellulose and
lignin in leaf litter. On an ecosystem basis, N additions decreased total annual litter
(leaf + fine root) fluxes of condensed tannins and soluble protein and increased the fluxes
of nonstructural carbohydrates (sugars + starch) and N (Xia etal.. 2015). Based on broad
decomposition chemistry patterns observed elsewhere, these changes in litter chemistry
were expected to increase initial rates of litter decomposition, which contrasts with the
decrease in litter turnover rates observed at these sites (Xia et al.. 2015). However,
changes in tissue chemistry are not ubiquitous; for example, Gricpcntrog et al. (2015)
observed that the abundance and composition of fatty acids in leaf and root tissues in
spruce and beech trees in Switzerland were not influenced by added N (70 kg N/ha/yr).
Further, while N additions can change plant tissue chemistry, links between litter
chemistry and microbial abundance and microbial function are complex [e.g., (Baumann
etal.. 2009)1.
6.2.3.3. Forest Lichens
Lichens are widely used as indicators of N deposition impacts on ecosystems, particularly
in forests. However, lichens also are important for ecosystem function. For insects, birds,
and mammals, lichens represent camouflage, building materials for nests, and a source of
food (Brodo et al.. 2001). Lichens absorb N, sulfur (S), and other elements from
atmospheric deposition and throughfall, and lichens that host cyanobacteria can add
June 2018
6-36
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
significantly to ecosystem N inputs, providing N to other plants (Kobvlinski and Frcdccn.
2015). In addition, lichens also have a role in hydrologic cycling, have many traditional
human uses, and have high potential for pharmaceutical use (McCunc and Geiser. 1997).
Lichens are symbioses comprised of fungi (mycobiont) and a green alga and/or
cyanobacterium [photobiont; (Sundbcrg etal.. 2001; Palmqvist. 2000)1. Much of the
lichen biomass is comprised of fungal (mycobiont) tissue, but the photobiont synthesizes
organic compounds, supplying the energy and structural C needed for growth of the
lichen. Both the photobiont and the mycobiont require N for growth (Palmqvist. 2000).
but the supply of C and N must be coordinated for the development of lichen thalli
(Sundbcrg et al.. 2001). Lichens with a cyanobacterial photobiont are N fixing, but those
with a green algal photobiont depend on atmospheric deposition for N.
Lichens can be classified based on their response to N pollution. Lichens occurring in
areas that receive high N deposition are considered nitrophytic or eutrophic; lichens
common in areas receiving low N input are designated acidophytic or oligotrophic (Gaio-
Oliveira et al.. 2005; van Herk. 2001). Atmospheric N deposition can impact lichens
through changes in physiological function caused by an increased supply of N or by
altering the pH of tree bark hosting the lichen (Jovan. 2008). Although lichens with a
green algal photobiont depend on atmospheric deposition as a source of N, these lichens
can also be negatively impacted by N deposition via the accumulation of toxic
concentrations of NH4 within the thallus. Cyanobacteria can grow on either NOs or
NH4+ sources when administered at nontoxic concentrations, but more rapid growth has
been observed with NH4 than NOs, (Von Ruckert and Giani. 2004). Ammonium is more
easily assimilated by lichens; both NOs and nitrite must first be reduced to NH4 before
assimilation can occur (Von Ruckert and Giani. 2004). The 2008 ISA noted that lichens
with a cyanobacterial photobiont appear to be more sensitive to adverse effects from
atmospheric N deposition than most other lichens (Dahlman et al.. 2002; Hallingback and
Kellner. 1992; Hallingback. 1991). In part, the sensitivity of lichens to increasing N
deposition is a function of the mechanisms with which that lichen can respond to high N
supply, such as decreasing N uptake or assimilating N into nontoxic forms such as
arginine (Gaio-Oliveira et al.. 2005; Dahlman et al.. 2002).
Since the 2008 ISA, new research on the impact of N deposition on the growth and
physiology of forest lichens has both confirmed that lichen abundance is sensitive to N
deposition and provided further insight on the growth and physiological changes that
occur when lichens are exposed to exogenous N (Table 6-5). For example, previous
research had suggested that lichens were most sensitive to N as ammonia (Sheppard et
al.. 2011; Jovan. 2008). However, a study of lichen communities on California black oak
(Quercus kelloggii) forests at 22 sites in the Los Angeles Basin in California, Jovan et al.
(2012) found that the abundance of eutrophic lichen species was only weakly related to
6-37
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
gaseous NH3 concentrations. Instead, the strongest N pollution-related predictor of
eutrophic lichen abundance was total N deposition (as canopy throughfall). Further, at the
relatively neutral bark pHs in the Los Angeles Basin, there was no influence of pH on the
abundance of eutrophic lichen species. The abundance of eutrophic lichens was also best
correlated with total throughfall N deposition in southeastern Alaska (Schirokaucr ct al..
2014a). This research provides strong evidence that total N deposition, not the deposition
of a particular form of N, is the primary driver of changes in the growth, physiology, and
composition of epiphytic lichens.
Increases in lichen thalli N concentrations in response to added N have been widely
observed in the U.S. and Europe (Table 6-5). even at relatively low rates of atmospheric
N deposition. For instance, McMurrav et al. (2013) measured throughfall N deposition
and sampled lichen thalli N concentrations at sites near the Bridger-Teton National Forest
Wilderness that were at increasing distances downwind of a major oil and natural gas
production field. Although the observed rates of N deposition were only 0.8 to
4.1 kg N/ha/yr along this gradient, thalli N concentrations in Usnea lapponica
approximately doubled from -1.2 to 2.4%. In southeastern Alaska, Schirokauer et al.
(2014a) found increases in thalli N concentrations along an even smaller range of N
deposition (0.05 to 1.05 kg N/ha/yr), a pollution gradient attributed to cruise ship
emissions.
In southern California, Riddell et al. (2008) transplanted thalli from the oligotrophic
lichen Ramalina menziesii from relatively unpolluted sites into fumigation chambers
under moderate and high HNO3 concentrations (19.9-25 (ig/m3 and 26.4-35.3 (ig/m3,
respectively). The HNO3 fumigation caused significant declines in chlorophyll content
and C exchange capacity compared to thalli in control chambers. This research was later
expanded to six species known to vary in sensitivity to N pollution (Riddell et al.. 2012).
Fumigation with HNO3 (daily peaks near 50 ppb) decreased chlorophyll content,
chlorophyll fluorescence, gross photosynthesis, and dark respiration in three of the five N
sensitive species; while only photosynthesis declined in the other two N sensitive species.
Four of the N sensitive species were tested for fumigation effects on cell membrane ion
leakage; overall ion leakage, and specifically K+ ion leakage, were increased by HNO3
fumigation in all species.
6-38
DRAFT: Do Not Cite or Quote

-------
Table 6-5 Growth and physiology responses of forest epiphytic lichens to
nitrogen added via atmospheric deposition or experimental
treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Nvbakken et al.
(2009)
Sweden and
Norway
Boreal
forests
Addition
50
1
C-based
secondary
compounds
Three
species: not
significant
One species:
decrease
Riddell et al.
(2008)
California
(Los Angeles
basin)
Oak forests
(Quercus
douglasii)
HNO3 gas
fumigation
0, 15,
30 mg/m3
0.08
Chlorophyll
content
Decrease
Nvbakken et al.
(2009)
Sweden and
Norway
Boreal forest
Addition
50
1
Chlorophyll
content
Three
species:
increase
One species:
not significant
Johansson et
al. (2010)
Sweden
Norway
spruce
{Picea
abies)
Addition
6, 12.5,
25, 50
3
Chlorophyll
content
Increase
Johansson et
al. (2011)
Sweden
Boreal forest
Addition
300
1
Chlorophyll
content
Increase
Asplund et al.
(2010)
Sweden and
Norway
Boreal
forests
Addition
50
1
Gastropod
feeding
preference
Three of four
lichen
species:
decrease
One species:
increase
Strenqbom and
Nordin (2008)
Sweden
Boreal forest
Addition
150
(twice)
Additions
22 and
30 yr
prior to
surveys
Lichen
abundance
Decrease
Johansson et
al. (2012)
Sweden
Norway
spruce
(Picea
abies)
Addition
6, 12.5,
25, 50
4
Lichen
abundance
Six and 12.5
doses:
increase
25 and 50
doses:
decrease
Will-Wolf et al.
(2015)
Northeastern
U.S.
Forests
Ambient
Not
stated
n/a
Lichen
abundance
Decrease
June 2018
6-39
DRAFT: Do Not Cite or Quote

-------
Table 6-5 (Continued): Growth and physiology responses of forest epiphytic
lichens to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Riddell et al.
(2008)
California
(Los Angeles
Basin)
Oak forests
(Quercus
douglasii)
HNO3 gas
fumigation
0, 15,
30 mg/m3
0.08
Membrane ion
leakage
Increase
Johansson et
al. (2011)
Sweden
Boreal forest
Addition
300
1
Mycobiont
growth
One species:
decrease
Two species:
not significant
Johansson et
al. (2011)
Sweden
Boreal forest
Addition
300
1
Photobiont
growth
Increase
Riddell et al.
(2008)
California
(Los Angeles
Basin)
Oak forests
(Quercus
douglasii)
HNO3 gas
fumigation
0, 15,
30 mg/m3
0.08
Photosynthesis
Decrease
Johansson et
al. (2011)
Sweden
Boreal forest
Addition
300
1
Photosynthesis
Increase
Riddell et al.
(2008)
California
(Los Angeles
Basin)
Oak forests
(Quercus
douglasii)
HNO3 gas
fumigation
0, 15,
30 mg/m3
0.08
Respiration
Decrease
Johansson et
al. (2010)
Sweden
Norway
spruce
{Picea
abies)
Addition
6, 12.5,
25, 50
3
Thalli N %
Six and
12.5 doses:
not significant
25 and
50 doses:
increase
Johansson et
al. (2011)
Sweden
Boreal forest
Addition
300
1
Thalli N %
Increase
McMurrav et al.
(2013)
Wyoming
Conifer
forests
Ambient
0.8-4.1
n/a
Thalli N %
Increase
Root et al.
(2013)
Western
North
America
Forests
Ambient
0.1-39.3
n/a
Thalli N %
Increase
Boltersdorf et
al. (2014)
Germany
Forests
Ambient
2.2-9.5
n/a
Thalli N %
Increase
Schirokauer et
al. (2014a)
Alaska
(southeast)
Conifer
forests
Ambient
0.05-1.05
n/a
Thalli N %
Increase
McMurrav et al.
(2015)
Idaho,
Wyoming,
Montana
Conifer
forests
Ambient
0.5-4.3
n/a
Thalli N %
Increase
June 2018
6-40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
Table 6-5 (Continued): Growth and physiology responses of forest epiphytic
lichens to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Johansson et
al. (2010)
Sweden
Norway
spruce
{Picea
abies)
Addition
6, 12.5,
25, 50
3
Thalli P %
Not significant
Johansson et
al. (2011)
Sweden
Boreal forest
Addition
300
1
Thalli P %
Two species:
increase
One species:
not significant
Nvbakken et al.
(2009)
Sweden and
Norway
Boreal forest
Addition
50
1
Thallus growth
Three
species:
increase
One species:
not significant
Johansson et
al. (2011)
Sweden
Boreal forest
Addition
300
1
Thallus growth
One species:
increase
One species:
not significant
One species:
decrease
C = carbon; ha = hectare; HN03 = nitric acid; kg = kilogram; m = meter; mg = milligram; N = nitrogen; n/a = not applicable;
P = phosphorus; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References are ordered by endpoint. Only
statistically significant effects are listed as increases or decreases.
Johansson and colleagues conducted a series of experiments in Sweden intended to
understand the physiological responses of epiphytic lichens to added N (Johansson et al..
2012; Johansson et al.. 2011; Johansson et al.. 2010). In a whole-tree N addition
experiment in a spruce forest in boreal Sweden, low rates of N addition (6,
12.5 kg N/ha/yr for 3 years) increased total lichen abundance, but higher N addition rates
(25, 50 kg N/ha/yr) decreased total lichen abundance (Johansson et al.. 2012). However,
there was considerable variation among lichen species in the N addition treatment rate
that resulted in optimal growth. For two of the lichen species in that experiment,
Alectoria sarmentosa and Platismatia glauca, Johansson et al. (2010) quantified
physiological responses to N additions. Thalli N concentrations increased in both species
at the two highest N addition rates (25, 50 kg N/ha/yr). There was a significant positive
relationship between cumulative N dose and chlorophyll content, but the N additions did
not affect thalli P concentrations. In a separate experiment with a single much higher N
addition rate (300 kg N/ha/yr for one season), Johansson et al. (2011) examined the
June 2018
6-41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
physiological and growth responses of three lichen species. All three lichen species
exhibited increased tissue N concentrations, increased chlorophyll concentrations,
increased photosynthesis, and increased growth of the photobiont. However, mycobiont
growth decreased in two species and was unchanged in the third. Total biomass changes
among the three species were positive, neutral, and negative.
In another experiment in Scandinavia, lichens collected from Sweden and Norway were
exposed to added N (50 kg N/ha/yr for one season) to understand whether this addition
altered the concentrations of the C based secondary compounds (CBSCs) thought to
protect lichens from herbivores and whether it affected feeding preferences of gastropod
herbivores (Asplund et al.. 2010; Nvbakken et al.. 2009). The N additions decreased the
concentrations of CBSCs in one species, but had no effect on the other three lichen
species (Nvbakken et al.. 2009). The gastropod herbivores preferred to feed on lichens
from the control treatment for three of the four species, while preferring the thalli from
the N addition treatment for the fourth lichen species (Asplund et al.. 2010). Notably, the
species exhibiting the decrease in CBSCs was not the species that was more preferable to
herbivores in the N addition treatment. Together, these results suggest that N deposition
may alter lichen community composition by shifting herbivore feeding preferences.
6.2.3.4. Net Ecosystem Production and Carbon Sequestration Response
The 2008 ISA concluded that N deposited onto terrestrial ecosystems increased net
primary productivity (NPP) and ecosystem C storage (kg C/ha). However, the extent to
which this additional N stimulates forest C sequestration was atopic of debate [e.g., (De
Schriiver et al.. 2008; de Vries et al.. 2008; Sutton et al.. 2008; Magnani et al.. 2007;
Nadelhoffer et al.. 1999b) I. Magnani et al. (2007) correlated net ecosystem production
(NEP; kg C/ha/yr) with estimated rates of N deposition for 20 forested sites in North
America and Europe, which resulted in an estimate of 725 kg C sequestered per kg of
added N (i.e., kg C/kg N). However, this estimate was widely criticized and a reanalysis
of these data produced an estimate of C sequestration more than an order of magnitude
lower [68 kg C/kg N; (Sutton et al.. 2008)1.
Since the 2008 ISA, several new syntheses and a number of field experiments and
modeling studies have provided further evidence that N deposition increases NPP, NEP,
and ecosystem C content, and that have more tightly constrained estimates of the
response of plant, soil, and ecosystem C content to N deposition. Nave et al. (2009b)
estimated that up to 15% of the N needed to support NPP in a northern Michigan aspen
forest was supplied by atmospheric N deposition. In China, N additions of 20, 50, or
100 kg N/ha/yr to an old growth boreal forest stimulated aboveground tree productivity
June 2018
6-42
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
by 5, 7, and 23% during the second and third years of the experiment (Du and Fang.
2014). Chen etal. (2011) found that a Douglas-fir stand in the Pacific Northwest
increased NEP by 2,500 kg/ha (+83%) in the first year after receiving 200 kg N/ha of
urea, both from decreased respiration (930 kg/ha) and increased gross primary production
(157 kg/ha). Increases in ecosystem C content have been noted in long-term N addition
experiments in Massachusetts (Trev et al.. 2014). Michigan (Prcgitzcr et al.. 2008). and
New York (Lovett et al.. 2013). Using a process model that neglected changes in stand
ontogeny, disturbances, and shifts in forest management, de Vries and Posch (2011)
concluded that N deposition was a dominant determinant of European forest productivity
throughout the 20th century.
Since the 2008 ISA, a variety of techniques have been used to quantify the C
sequestration response of forests to N deposition, particularly in Europe (Frev et al..
2014). Using a 15N-labeling technique, Gundale et al. (2014) observed a linear
relationship between C sequestration and N additions of up to 50 kg N/ha/yr that had a
slope of 16 kg C/kg N in a Swedish boreal forest. Hvvonen et al. (2008) synthesized soil
and plant C sequestration data from 15 long-term (14-30 year) N addition experiments in
boreal (Picea abies, Pinus sylvestris) forests in Sweden and Finland and estimated that C
sequestration averaged 23 kg C/kg N for Picea and 30 kg C/kg N for Pinus, with an
additional 11 kg C/kg N within the soil. Also in Sweden, Eliasson and Agren (2011)
applied an ecosystem model to Scots pine (Pinus sylvestris) forests and estimated that
ecosystem C stocks increased by 24,123 kg/ha in response to a cumulative 224 kg N/ha
added over a century (108 kg C/kg N). Based on the stimulation of photosynthesis and
assumptions about the fraction of photosynthate that is ultimately transformed into tree
biomass, Fleischer et al. (2013) estimated that N deposition stimulates C sequestration by
25 kg C/kg N.
Forest inventory studies typically make use of one or more decades of tree growth data
from national or continental monitoring networks and relate the variation in this growth
to differences in N deposition and other environmental factors. Across western and
northern Europe, Solberg et al. (2009) found that variation in tree volume increment was
positively related to N deposition and summer temperature, particularly for pine (Pinus)
and spruce (Picea), with similar, but weaker, relationships for beech (Fagus) and oak
(Quercus). The overall model estimated C sequestration was 19 kg C/kg N. Eastaugh et
al. (2011) analyzed Norway spruce (Picea abies) growth data from the Austrian National
Forest Inventory and estimated thatN deposition sequestered 21.6 kg C/kg N in
aboveground tree biomass during the latter half of the 21st century. Etzold et al. (2014)
and Ferretti et al. (2014) took similar approaches with 18 inventory plots in Switzerland
and 25 inventory plots in Italy, respectively. Both studies found a positive relationship
between N deposition and NPP, but were unable to isolate NPP effects from other
6-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
environmental factors. Within the U.S., the analysis of forest inventory data in the
northeastern U.S. by Thomas et al. (2010) estimated that N deposition increased
aboveground tree biomass at 61 kg C/kg N; a rough estimate including belowground C
increased this to 73 kg C/kg N. Pinder et al. (2012) modified the estimates of Thomas et
al. (2010). using alternate values for N deposition, more complex belowground biomass
estimates, and changes in soil C pools, to produce an enhancement of 65 kg C/kg N. One
criticism of the inventory and modeling studies is that they match variation in growth
only to current levels of N deposition and neglect the potential effects of previous N
deposition at that site. Thus, the influence of current N deposition may be exaggerated
(Hogberg. 2012). In addition, estimates of tree growth and biomass production typically
rely on the application of equations that predict tree biomass based on measurements of
tree stem diameter, either alone or in combination with tree height measurements. Ibanez
et al. (2016) observed that at the sites in Michigan where N additions had increased tree
growth (Pregitzer et al.. 2008). the N additions had also altered the allometric patterns of
sugar maple tree growth so that at a given tree stem diameter at breast height (1.4 m), the
trees receiving N additions were taller than those in the control plots. This meant that
stemwood biomass equations underestimated tree size for trees receiving N additions
(Ibanez et al.. 2016). If these results apply in other environments and with other tree
species, this implies that allometric biomass equations may be underestimating tree
biomass in regions experiencing high rates of atmospheric N deposition and that
estimates of C sequestration due to tree growth increases [e.g., (Thomas et al.. 2010)1
may be too low.
In the long-term N addition experiment at Harvard Forest (Frev et al.. 2014). the C
response to N deposition ranged from -2 to 30 kg C/kg N among the four treatment types
(50 or 150 kg N/ha/yr; Pinus or Quercus forests) and the amount of C sequestered in the
soil was equal to or greater than the amount of C sequestered in trees. Similar results
showing more C sequestration in the soil than in trees were found in N addition
experiments at four Michigan forests (23 kg C/kg N soil vs. 17 kg C/kg N tree; (Pregitzer
et al.. 2008) and in the Catskills (Lovettetal.. 2013). Several syntheses of forest C
sequestration changes in response to N deposition have been conducted (Tian et al..
2016a; Frev et al.. 2014; Pinder etal.. 2013; Pinder etal.. 2012; Butterbach-Bahl et al..
2011; Janssens et al.. 2010; Liu and Greaver. 2009). producing estimates of 12-41 kg
C/kg N. Often these synthesis studies are meta-analyses; however, de Vries et al. (2014a)
took a novel approach of combining estimates of ecosystem N retention with N allocation
and C:N ratios to develop stoichiometric estimates ofN deposition for tropical,
temperate, and boreal ecosystems. Although this approach involves a number of
important assumptions about biogeochemistry and tree physiology, estimates for tree and
soil C sequestration response rates were similar to values from other syntheses
(Figure 6-3). In their study of aboveground NPP responses per unit of added N, Tian et
6-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
al. (2016a) found that the average forest response was a 3.75% increase in NPP per g of
N.
Variation in these estimates is likely influenced by the research approach. For instance,
the model used by Eliasson and Agren (2011) predicted almost no N losses from the
forests due to leaching or denitrification, which are both widely reported at high rates of
N deposition. However, environmental and ontological factors also appear to influence
the sensitivity of forest C sequestration to N deposition. Using the G'DAY model, Dezi et
al. (2010) found recent forest harvests, shifts in plant C allocation away from roots, and
canopy N uptake all increased the C sequestration response to N deposition. Among the
15 experiments synthesized by Hwonen et al. (2008). the amount of C sequestered
ranged from -0.8 to 61 kg C/kg N, with greater increases in kg C/kg N at lower N doses.
The amount of additional tree C sequestered was small in plots that had low soil
O-horizon C:N ratios (near 25) and the amount of C sequestered by added N increased
with O-horizon C:N until C:N reached 35. Among Picea forests, young stands were more
responsive than old stands, but no such effect was apparent for Pinus. In addition, C
sequestration increased more in plots simultaneously supplied with K and P (Hwonen et
al.. 2008). Janssens et al. (2010) acknowledged the role of stand age in altering the
dynamic of C cycling, explicitly removing young, rapidly expanding forests from some
portions of the meta-analysis.
6-45
DRAFT: Do Not Cite or Quote

-------
Aboveground
¦ 9
:orest Biomass
dD o
-O-
Modeling
J*
^ 0
. oO o
oo oaO c
t&o
Inventory
Experimental
°o
Fl«ischct etal. 2011
deVneietil. 2006
Thomas etai. 2010
ŁiiuuŁhetU. 2011
iolberjml. 2009
Laubhann el al, 2009
Wade lhcfler«»l. 1999
Hojtefj et >l. 2006
H>vonc-nfUl.2COS
Pr*tit»r«»l.2006
L overt «l if. 2013
Frvy et al. 2014
GumUfetft *1.2014
Fowler et al. 20 IS
~O
A.
Synthesis
•30	0	30	60
Response Ratio (kg C kf1 N)
Ecosystem Carbon O
nO o
•—O—
Modeling
ButtMbKh^ah! at it. 2011
daVrinwtal.2014
da VtsH at al. 2014
de Vrsn et al. 2014
Levy et al- 200J
Sunonef»U2008
Sutton «t »1.2008[one sfctji
Wamelink etal- 2009b
Warns link «t al. 2009a
lliasson & Agren 2011
Inventory
tf b VriM II al. 20K
Sutton at al. 2QQ6[Map»ni
> oO
0
o
oo O o
O o
O	0
o _0
o
0 o
Experimental

¦o

Synthesis
N*d*lh
-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Tundra areas in the U.S. are concentrated in high-elevation alpine areas in the western
U.S. and high-latitude areas in Alaska. Although alpine ecosystems are limited in their
spatial extent, these ecosystems are important components of many national parks
(e.g., Rocky Mountain, Yosemite, Sequoia-Kings Canyon, Mount Rainier) and other
Class I areas within these regions. Nitrogen deposition rates in Arctic tundra areas are
generally low, but there is a distinct impact of anthropogenic pollution on N cycling that
could increase with further industrial development at high latitudes (Holtgricvc et al..
2011). In addition, although baseline rates of N deposition in Arctic deposition tend to be
low, isolated precipitation events carrying air masses from industrialized or agricultural
regions can result in high rates of deposition (Kuhncl et al.. 201IV In Svalbard (north of
Norway), 10% of precipitation events were responsible for 93% of all N deposited in
snow or rain (Kuhncl etal.. 2011).
For Arctic tundra ecosystems, the 2008 ISA identified a single N addition experiment and
cited two studies from the 1980s on the effects of decreased N availability. The two
studies on plants grown under conditions of low N availability found that tundra plants
responded by increasing root growth (Bloom etal.. 1985) or increasing N use efficiency
(Chapin. 1980). The N addition (50 or 100 kg N/ha/yr) experiment began in 1981 and
caused changes in plant community composition (Shaver et al.. 2001) and increases in
aboveground plant growth, but losses in soil C pools that resulted in decreased ecosystem
C content (Mack et al.. 2004).
The 2008 ISA reported that alpine plant species are often sensitive to N enrichment
because many are adapted to low nutrient availability (Bowman et al.. 2006) and that N
deposition may increase alpine plant productivity and alter plant community composition
(Neff et al.. 2002; Seastedt and Vaccaro. 2001; Bowman etal.. 1995). Some alpine plants
in the southern Rocky Mountains exhibit greater growth in response to increased N, but
this species-specific growth response to N deposition is one of the mechanisms that can
change community composition (Bowman et al.. 1993). Many of the dominant plant
species do not respond to additional N supply with increased production. Rather, many
subdominant species increase in abundance when the N supply is increased (Fenn et al..
2003a). The 2008 ISA also noted that the effects of N deposition can be spatially
heterogeneous in alpine tundra ecosystems, with areas that accumulate wind-blown snow
receiving higher rates of winter N deposition (Bowman and Steltzer. 1998). Further, the
2008 ISA also cited evidence that chronic N additions to alpine ecosystems can
accelerate the decomposition of some soil C fractions, while preserving other C fractions
(Neffetal.. 2002).
Research since 2008 on the effects of added N on alpine and Arctic ecosystems has
included broad syntheses, a number of experiments in both Europe and Asia, and
6-47
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
additional research in North America. Bouskill et al. (2014) conducted a meta-analysis of
high-latitude N addition experiments and compared these results with model simulations
from the Community Land Model (CLM), the land component of the Community ESM
(CESM). Bouskill et al. (2014) identified 37 N addition field experiments from 14 sites in
North America and Europe. In this synthesis, N additions significantly stimulated gross
primary production (GPP; +44 ± 7%, mean ± standard error), while soil respiration
declined by 12 ± 7%. Total aboveground plant biomass was not significantly affected by
N additions (increase of 15 ± 22%), but vascular plant biomass increased by 33 ± 8%.
Notably, the modeled responses produced from CLM matched observations only for two
parameters: soil organic matter and GPP. These matches occurred only under the lowest
rates of N addition (<1 kg N/ha/yr), suggesting that N cycling processes or parameters are
not well characterized in tundra ecosystems.
Three multiyear experiments span the period before and after the 2008 ISA was
published (Table 6-6). In a subalpine shrub heathland in Scotland, Britton and Fisher
(2007) examined the effects on plant communities of burning, clipping, and N additions
of 0, 10, 20, or 50 kg N/ha/yr as NH4NO3 for 5 years. Since the 2008 ISA was published,
Britton et al. (2008) observed that N additions increased shoot N concentration in the
dominant shrub Calluna vulgaris, with significant increases only at the two highest levels
of N addition. In assessing plant growth, Britton and Fisher (2008) found increased
growth of the dominant alpine shrub Calluna vulgaris and of Vaccinium vitis-idaea with
N additions of 20 and 50 kg N/ha/yr, but no effect with 10 kg N/ha/yr or with N at any
level on three other shrub species. Notably, in a greenhouse study with Calluna vulgaris
plants from four different geographic populations, Mever-Gruenefeldt et al. (2016)
observed that N additions (35 kg N/ha/yr for 2 years) increased aboveground biomass in
all populations, but that responses varied by factor of two between populations.
Belowground biomass was unresponsive to N addition in the greenhouse study, resulting
in decreased root: shoot ratios and potentially making the plants more susceptible to
drought. In a second study within Scotland, Britton and Fisher (2010) found that N
additions of 7.5, 12.5, and 22.5 kg N/ha/yr increased thallus N concentration in three of
four alpine lichen species, but decreased growth of two species and did not significantly
affect growth in three other species.
6-48
DRAFT: Do Not Cite or Quote

-------
Table 6-6 Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added via
atmospheric deposition or experimental treatments.
Ambient	Nitrogen	Effect of
Deposition or Addition Rate Duration	Additional
Reference	Study Location Vegetation	Addition	kg N/ha/yr	yr	Endpoint	Nitrogen
Kellev and Epstein (2009) Alaska
Tundra meadow Addition
(Dryas integrifolia,
Eriophorum
vaginatum, Carex
spp.)
100
Above-ground
plant biomass
Not significant
Volketal. (2011)
Switzerland Subalpine	Addition
grassland (Nardus
stricta, Carex
sempervirens,
Festuca spp.)
5, 10, 25, 50
Above-ground
plant biomass
Low dose: not
significant
Other doses:
increase
Bassin et al. (2012)
Switzerland Subalpine	Addition
grassland (Carex
sempervirens)
50
Above-ground
plant biomass
Increase
Blanke et al. (2012)	Switzerland Subalpine	Addition	50	3 Above-ground Increase
grassland (Festuca	plant biomass
rubra, F. violacea,
Nardus stricta,
Carex
sempervirens)
Blanke etal. (2012)
Switzerland Subalpine	Addition
grassland (Festuca
violacea,
Leontodon
helveticus, Carex
sempervirens,
Trifolium alpinum)
50
Above-ground
plant biomass
Festuca'. increase
Leontodon, Carex,
Trifolium'. not
significant
June 2018
6-49
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
kg N/ha/yr
Duration
yr
Endpoint
Effect of
Additional
Nitrogen
Bowman et al. (2012)
Colorado (Rocky
Mountain
National. Park)
Dry sedge
meadow (Kobresia
myosuroides,
Carex rupestris)
Addition
5, 10, 30
4
Above-ground
plant biomass
Not significant
Onipchenko et al. (2012)
Russia
(Caucasus
Mountains)
Lichen heath
(Cetraria islandica)
Addition
90
5
Above-ground
plant biomass
Increase
OniDchenko et al. (2012)
Russia
(Caucasus
Mountains)
Alpine grassland
(Festuca varia)
Addition
90
5
Above-ground
plant biomass
Not significant
OniDchenko et al. (2012)
Russia
(Caucasus
Mountains)
Alpine meadow
(Geranium
gymnocaulon)
Addition
90
5
Above-ground
plant biomass
Not significant
Onipchenko et al. (2012)
Russia
(Caucasus
Mountains)
Alpine snowbeds
(Sibbaldia
procumbens)
Addition
90
5
Above-ground
plant biomass
Not significant
Bouskill et al. (2014)
North America
and Europe
Arctic and high
latitude
Addition
Average: 72
range: 1-100
Meta-
analysis
Above-ground
plant biomass
Not significant
Gill (2014)
Utah
Subalpine meadow Addition
(Achnatherum
lettermanii,
Artemisia
michauxiana)
70
3
Above-ground
plant biomass
Increase
Volketal. (2014)
Switzerland
Subalpine
grassland (Nardus
stricta, Carex
sempervirens,
Festuca spp.)
Addition
5, 10, 25, 50
7
Above-ground
plant biomass
Increase
June 2018
6-50
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.



Ambient
Nitrogen


Effect of



Deposition or
Addition Rate
Duration

Additional
Reference
Study Location
Vegetation
Addition
kg N/ha/yr
yr
Endpoint
Nitrogen
Zamin et al. (2014)
Northwest
Shrub tundra
Addition
10, 100
8
Above-ground
Low dose: not

Territories
(Vaccinium



plant biomass
significant

(Canada)
vitisidaea,




Hiqh dose:


Rhododendron




decrease


subarcticum,







Andromeda







polifolia)





Farrer et al. (2015)
Colorado (Niwot
Moist alpine
Addition
229
7
Above-ground
Deschampsia:

Ridge)
meadow



plant biomass
increase


(Deschampsia




Geum\ decrease


cespitosa, Geum







rossii)





Sona and Yu (2015)
China (Tibetan
Alpine meadow
Addition
3.75, 15, 75
8
Above-ground
Low and mid dose:

Plateau)
(Kobresia humilis,



plant biomass
not significant


Elymus nutans,




Hiqh dose: increase


Stipa aliena,






Festuca ovina)





Blanke etal. (2012)
Switzerland
Subalpine
Addition
50
3
Above-ground
Grasses: increase


grassland (Festuca



plant biomass
Forbs, sedqes.


rubra, F. violacea,



(functional group)
lequmes: not


Nardus stricta,




significant


Carex






sempervirens)





Bassin et al. (2013)
Switzerland
Subalpine
Addition
5, 10, 25, 50
7
Above-ground
Sedqe, qrass:


grassland (Nardus



plant biomass
increase


stricta, Carex



(functional group)
Forbs. lequmes: not


sempervirens,




significant


Arnica montana,






Gentiana acaulis)





Bouskill etal. (2014)
North America
Arctic and high
Addition
Average: 72
Meta-
Above-ground
Increase

and Europe
latitude

range: 1-100
analysis
plant biomass







(vascular)

June 2018
6-51
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
kg N/ha/yr
Duration
yr
Endpoint
Effect of
Additional
Nitrogen
Zamin and Groqan (2012)
Northwest
Territories
(Canada)
Shrub tundra
(Betula
glandulosa,
Vaccinium
vitisidaea,
Rhododendron
subarcticum)
Addition
10, 100
7
Above-ground
plant growth
(Betula)
Low dose: not
significant
Hiqh dose: increase
Blanke et al. (2012)	Switzerland Subalpine	Addition	50	1	Below-ground
grassland (Festuca	plant biomass
violacea,
Leontodon
helveticus, Carex
sempervirens,
Trifolium alpinum)
Volk et al. (2014)	Switzerland Subalpine	Addition	10, 50	7	Below-ground Increase
grassland (Nardus	plant biomass
stricta, Carex
sempervirens,
Festuca spp.)
Arens et al. (2008)	Greenland	Dwarf-shrub/herb Addition
tundra (Salix
arctica, Carex
rupestris, Dryas
integrifolia)
Armitaae et al. (2011) U.K. Bryophyte	Ambient Reciprocal	2 Bryophyte Decrease
heathlands	transplant (7.2	biomass (R.
(Ranicomitrium	with 8.2-32.9)	lanuginosum)
lanuginoseum)
Festuca: increase
Leontodorr.
decrease
Carex. Trifolium'. not
significant
5, 10, 50	3	Below-ground Low and high dose:
respiration	not significant
Mid dose: increase
June 2018
6-52
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
kg N/ha/yr
Duration
yr
Endpoint
Effect of
Additional
Nitrogen
Wardleet al. (2013)
Sweden
Tundra meadow
(Deschampsia
flexuosa,
Empetrum
hermaphroditum,
Vaccinium spp.)
Addition
50
21
Bryophyte cover
Decrease
Armitaqe et al. (2012)
Europe (North
Atlantic)
Alpine heathlands
(Ranicomitrium
lanuginoseum)
Ambient
0.6-39.6
n/a
Bryophyte cover
(R. ianuginosum)
Decrease
Armitaqe et al. (2011)
U.K.
Bryophyte
heathlands
(Ranicomitrium
lanuginoseum)
Ambient
Reciprocal
transplant (7.2
with 8.2-32.9)
2
Bryophyte growth
(R. ianuginosum)
Increase
Armitaae et al. (2012)
Europe (North
Atlantic)
Alpine heathlands
(Ranicomitrium
ianuginosum)
Ambient
0.6-39.6
n/a
Bryophyte growth
(R. ianuginosum)
Increase
Armitaae et al. (2011)
U.K.
Bryophyte
heathlands
(Ranicomitrium
ianuginosum)
Ambient
Reciprocal
transplant (7.2
with 8.2-32.9)
2
Bryophyte tissue N
% (R.
ianuginosum)
Increase
Armitaae et al. (2012)
Europe (North
Atlantic)
Alpine heathlands
(Ranicomitrium
ianuginosum)
Ambient
0.6-39.6
n/a
Bryophyte tissue N
% (R.
ianuginosum)
Increase
Armitaae et al. (2012)
Europe (North
Atlantic)
Alpine heathlands
(Ranicomitrium
ianuginosum)
Ambient
0.6-39.6
n/a
Bryophyte tissue P
% (R.
ianuginosum)
Increase
Wana et al. (2017a)
China (Tibetan
Plateau)
Alpine shrubland
(Sibiraea
angustata)
Addition
20, 50, 100
1.5
Ecosystem C
pools (shrubs,
grass, litter and
soil)
Increase; linear
increase with N
June 2018
6-53
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.



Ambient
Nitrogen


Effect of



Deposition or
Addition Rate
Duration

Additional
Reference
Study Location
Vegetation
Addition
kg N/ha/yr
yr
Endpoint
Nitrogen
Arens et al. (2008)
Greenland
Dwarf-shrub/herb
Addition
5, 10, 50
3
Ecosystem
Increase


tundra (Salix



respiration



arctica, Carex







rupestris, Dryas







integrifolia)





Volketal. (2011)
Switzerland
Subalpine
Addition
10, 50
4
Ecosystem
Not significant


grassland (Nardus



respiration



stricta, Carex







sempervirens,







Festuca spp.)





Liu et al. (2017b)
China (Tibetan
Alpine meadow
Addition
50
2
Flowering height of
Increase in

Plateau)




three perennial
flowering height in






forb species
two of the species;







not significant for







the remaining







species
Yuan et al. (2016)
Colorado (Niwot
Alpine meadow
Addition
50-200 (varied
20
Foliage biomass
Not significant for

Ridge)


over the 20 yr

(total-graminoid
total and forb only




duration;

and forb;
biomass; graminoid




averaged ca.

graminoid only;
biomass increased




85)

and forb only)
in moist and wet







meadow type, but







not dry meadow







type
Bassin et al. (2009)
Switzerland
Subalpine
Addition
50
3
Foliar N %
Six species:


grassland (Festuca




increase


rubra, Nardus




Two species: not


stricta, Carex




significant


sempervirens)




June 2018
6-54
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Ambient	Nitrogen	Effect of
Deposition or Addition Rate Duration	Additional
Reference	Study Location Vegetation	Addition	kg N/ha/yr	yr	Endpoint	Nitrogen
Kellev and Epstein (2009) Alaska
Tundra meadow Addition
(Dryas integrifolia,
Eriophorum
vaginatum, Carex
spp.)
100
Foliar N %
Increase
Bishop et al. (2010)
Washington
(Mount. St.
Helens)
Primary
successional
alpine meadow
(Agrostis pallens,
Lupinus lepidus)
Addition
78
4
Foliar N %
Aarostis'. increase
Lupinus'. decrease
Bowman et al. (2012)
Colorado (Rocky
Mountain
National Park)
Dry sedge
meadow (Kobresia
myosuroides,
Carex rupestris)
Addition
5, 10, 30
4
Foliar N %
Not significant
Bassin et al. (2012)
Switzerland
Subalpine
grassland (Carex
sempervirens)
Addition
50
2
Foliar N % (C.
sempervirens)
Increase
Britton et al. (2008)
Scotland
Shrub heathland
(Calluna vulgaris)
Addition
10, 20, 50
5
Foliar N %
(Calluna vulgaris)
Low dose: not
significant
Mid and hiah dose:
increase
Southon etal. (2013)
U.K.
Heathlands
(Calluna vulgaris)
Ambient
5.9-32.4
n/a
Foliar N %
(Calluna vulgaris)
Not significant
Blanke etal. (2012)
Switzerland Subalpine	Addition
grassland (Festuca
rubra, F. violacea,
Nardus stricta,
Carex
sempervirens)
50
Foliar N %	Grasses: increase
(functional group) Forbs. sedges: not
significant
June 2018
6-55
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
kg N/ha/yr
Duration
yr
Endpoint
Effect of
Additional
Nitrogen
Arens et al. (2008)
Greenland
Dwarf-shrub/herb
tundra (Salix
arctica, Carex
rupestris, Dryas
integrifolia)
Addition
5, 10, 50
3
Gross ecosystem
production
Increase
Volketal. (2011)
Switzerland
Subalpine
grassland (Nardus
stricta, Carex
sempervirens,
Festuca spp.)
Addition
10, 50
4
Gross primary
production
Not significant
Bouskill et al. (2014)
North America
and Europe
Arctic and high
latitude
Addition
Average: 72
range: 1-100
Meta-
analysis
Gross primary
productivity
Increase
Aerts (2009)
Sweden
Tundra meadow
(Betula nana)
Addition
75
10
Inflorescence
production
Decrease
Zamin and Groqan (2012)
Northwest
Territories
(Canada)
Shrub tundra
(Betula
glandulosa)
Addition
10, 100
7
Inflorescence
production
Not significant
Bassin et al. (2009)
Switzerland
Subalpine
grassland (Festuca
rubra, Nardus
stricta, Carex
sempervirens)
Addition
50
3
Leaf chlorophyll
concentration
Nine species:
increase
One species: not
significant
Bassin et al. (2012)
Switzerland
Subalpine
grassland (Carex
sempervirens)
Addition
50
2
Leaf length (C.
sempervirens)
Increase
Bassin et al. (2009)
Switzerland
Subalpine
grassland (Festuca
rubra, Nardus
stricta, Carex
sempervirens)
Addition
50
3
Leaf mass (per
leaf)
Three species:
increase
Seven species: not
significant
June 2018
6-56
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Ambient	Nitrogen	Effect of
Deposition or Addition Rate Duration	Additional
Reference	Study Location Vegetation	Addition	kg N/ha/yr	yr	Endpoint	Nitrogen
Aerts (2009)
Sweden	Tundra meadow
(Empetrum
hermaphroditum,
Andromeda
polifolia, Betula
nana, Eriophorum
vaginatum)
Addition
75
10
Leaf production Andromeda:
increase
Other three species:
not significant
Aerts (2009)
Sweden
Tundra meadow
(Empetrum
hermaphroditum,
Andromeda
polifolia,
Eriophorum
vaginatum)
Addition
75
10
Leaf survival
Andromeda and
Eriophorum'.
decrease
Empetrum'. not
significant
Arens et al. (2008)
Greenland
Dwarf-shrub/herb
tundra (Salix
arctica, Carex
rupestris, Dryas
integrifolia)
Addition
5, 10, 50
Net ecosystem
exchange
Low dose: not
significant
Mid and high dose:
decrease
Volketal. (2011)
Switzerland Subalpine	Addition
grassland (Nardus
stricta, Carex
sempervirens,
Festuca spp.)
10, 50
Net ecosystem
production
Low dose: not
significant
High dose:
decrease
Farrer et al. (2015)
Colorado (Niwot
Ridge)
Moist alpine
meadow
(Deschampsia
cespitosa, Geum
rossii)
Addition
229
Net primary
productivity
Not significant
June 2018
6-57
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Reference
Study Location Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
kg N/ha/yr
Duration
yr
Endpoint
Effect of
Additional
Nitrogen
Liu etal. (2017b)
China (Tibetan Alpine meadow
Plateau)
Addition
50
Phenology of three
perennial forb
species—first and
last flowering time
Delay in first
flowering time for
two of the species,
not significant for
the remaining
species; delay in
last fruiting time for
two of the species,
moved-up last
fruiting time in the
remaining species
Liu etal. (2017b)
China (Tibetan
Plateau)
Alpine meadow
Addition
50
Phenology of three
perennial forb
species—first and
last fruiting time
Delay in first fruiting
time for one of the
species, not
significant for the
others; delay in last
fruiting time for one
of the species, not
significant for the
others
Liu etal. (2017a)
China (Tibetan
Plateau)
Alpine meadow
Addition
50
1-2 Phenological traits
of six species (two
graminoids and
four forbs)
Change in fruiting
date in one of the
2 yr; not significant
for the remaining
traits (flowering
date, flowering
duration, fruiting
duration, and
growing duration) or
years
June 2018
6-58
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.



Ambient
Nitrogen


Effect of



Deposition or
Addition Rate
Duration

Additional
Reference
Study Location
Vegetation
Addition
kg N/ha/yr
yr
Endpoint
Nitrogen
Yin etal. (2017)
China (Tibetan
Alpine meadow
Addition
75
8
Phenological traits
Change depended

Plateau)




of six species (two
upon form of N






grasses, one
added and species;






sedge, and three
NH4+ modified






forbs)
reproductive







phenology of three







species; NO3"







delayed







senescence for







Elymus nutans;







effects of NH4NO3







were not significant
Field etal. (2017)
Wales
Shrub heathland
Addition
10, 20, 40, 120
10-25
Plant biomass C
Increase


(Calluna spp.)





Wanqetal. (2017a)
China (Tibetan
Alpine shrubland
Addition
20, 50, 100
1.5
Plant biomass
Increase; linear

Plateau)
(Sibiraea



(total shrub and
increase with N


angustata)



grass biomass)
mainly due to







increased root







biomass in shrubs







and grasses
Bishop et al. (2010)
Washington (Mt.
Primary
Addition
78
5
Plant cover
Increase

St. Helens)
successional







alpine meadow





Bishop et al. (2010)
Washington (Mt.
Primary
Addition
78
5
Plant cover
Forbs: increase

St. Helens)
successional



(functional group)
Graminoids: not


alpine meadow




significant







Leaumes: decrease
Armitaqe et al. (2014)
Europe (North
Alpine heathlands
Ambient
0.6-39.6
n/a
Plant cover
Shrubs: decrease

Atlantic)




(functional group)
Forbs: not







significant







Graminoids:







increase
June 2018
6-59
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
kg N/ha/yr
Duration
yr
Endpoint
Effect of
Additional
Nitrogen
Field etal. (2017)
Wales
Shrub heathland
(Calluna spp.)
Addition
10, 20, 40, 120
10-25
Plant height
Increase
Field etal. (2017)
Wales
Shrub heathland
(Calluna spp.)
Addition
10, 20, 40, 120
10-25
Plant I itte rf a 11
Increase
Farreretal. (2013)
Colorado (Niwot
Ridge)
Moist alpine
meadow (Geum
rossii)
Addition
288
11
Plant nonstructural
carbohydrate
pools
Decrease
Wanqetal. (2017a)
China (Tibetan
Plateau)
Alpine shrubland
(Sibiraea
angustata)
Addition
20, 50, 100
1.5
Plant rootshoot
ratio
Increase
Churchland et al. (2010)
Northwest
Territories
(Canada)
Shrub tundra
(Betula
glandulosa,
Vaccinium
vitisidaea,
Rhododendron
subarcticum)
Addition
100
1
Plant tissue N %
Increase
Britton and Fisher (2008)
Scotland
Shrub heathland
(Calluna vulgaris)
Addition
10, 20, 50
5
Shoot growth
(Calluna vulgaris)
Low dose: not
significant
Mid and hiah dose:
increase
Britton and Fisher (2008) Scotland	Shrub heathland Addition	10,20,50	5	Shoot growth Not significant
(Calluna vulgaris)	(Empetrum
hermaphroditum,
Arctostaphylos
uvaursi, Vaccinium
myrtillus)
June 2018
6-60
DRAFT: Do Not Cite or Quote

-------
Table 6 6 (Continued): Alpine and Arctic tundra plant productivity and physiology responses to nitrogen added
via atmospheric deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
kg N/ha/yr
Duration
yr
Endpoint
Effect of
Additional
Nitrogen
Britton and Fisher (2008)
Scotland
Shrub heathland
(Calluna vulgaris)
Addition
10, 20, 50
5
Shoot growth
(Vaccinium
vitisidaea)
Low dose: not
significant
Mid and hiqh dose:
increase
Bassin et al. (2009)
Switzerland
Subalpine
grassland (Festuca
rubra, Nardus
stricta, Carex
sempervirens)
Addition
50
3
Specific leaf area
One species:
increase
Two species:
decrease
Eiqht species: not
significant
Wardleet al. (2013)
Sweden
Tundra meadow
(Deschampsia
flexuosa,
Empetrum
hermaphroditum,
Vaccinium spp.)
Addition
50
21
Vascular plant
cover
Increase
ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; NEE = net ecosystem exchange; yr = year.
Notes: single studies are reported more than once if multiple endpoints or ecosystems were measured. References ordered by endpoint. Only statistically significant effects are listed
as increases or decreases.
June 2018
6-61
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
In subalpine tundra in Sweden, a long-term experiment was started in 1989 to understand
how alleviating nutrient limitations altered ecological processes, including productivity,
decomposition, and the development of plant communities ("Wardle et al.. 2013; Nilsson
ct al.. 2002). The experimental design included six treatments, including additions of
N03 and NH4NO3 at rates of 50 kg N/ha/yr. After the first 9 years, there were negative
effects of both NO3 and NH4NO3 on the percentage cover of the dominant ericaceous
shrub Empetrum hermaphroditum, but increased cover of two ericaceous Vaccinium
shrub species (Nilsson et al.. 2002). An even stronger positive response was exhibited by
the bunchgrass Deschampsia flexuosa, which became the dominant plant in plots
receiving N additions. The N additions also had varying effects on the cover of the
dominant bryophyte species. Dicranum spp. decreased in cover in response to added N
and Pleurozium schreberi disappeared entirely from some N addition plots, while Polia
spp., Bryum spp. and Barbilophozia lycopodioides were not significantly affected by N
additions. The dominant lichen, Cladina spp., decreased in response to added N (Nilsson
et al.. 2002). The N addition treatments did not impact vascular plant diversity, but moss
species richness was decreased by NH4NO3 and lichen species richness was decreased by
NC>3~. Decomposition of pine needles in litterbags was followed for 4 years, but no
significant effects of the N addition treatments were observed (Nilsson et al.. 2002). In
2010, Wardle et al. (2013) reassessed the changes in plant cover caused by the N addition
treatments. The N addition treatments increased cover of vascular plants, but decreased
bryophyte cover and lichen cover. Specifically, N additions repeatedly increased cover of
Deschampsia and decreased cover of Empetrum, although these changes were not always
significant. Within the soil microbial community, NH4NO3 additions decreased the
abundance of fungal phospholipid fatty acids (PLFAs; a marker of microbial abundance),
while bacterial PLFAs were increased by NO3 additions. However, NO3 and NH4NO3
each decreased the fungal-to-bacterial PLFA ratio in the soil. Humus mass and C content
were increased by NH4NO3 additions, but not significantly affected by NO3 .
Also spanning the period before and after the 2008 ISA was a 7-year N deposition study
in a subalpine grassland in the Swiss Alps, with rates of N addition of 0, 5, 10, 25, or
50 kg N/ha/yr (as NH4NO3). The initial analysis of the first 3 years of data found that N
additions increased aboveground plant productivity (Bassin et al.. 2007). Among
functional groups, sedge growth was consistently stimulated, forb growth increased
strongly in the first year of the experiment, and grasses and legumes were unresponsive
to N. Species richness was not influenced by the N additions, but N did alter the relative
abundance of the 11 most frequently occurring species. A subsequent analysis of this
experiment Bassin et al. (2009) focused on the physiological responses of these
11 species to the 50 kg N/ha/yr treatment. The N addition treatment increased leaf
concentrations of N and chlorophyll in 9 of the 11 species. Notably, the species
exhibiting the largest growth response (Carex sempervirens) also showed the largest
6-62
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
increase in foliar concentrations of N and chlorophyll. Four years into the experiment,
Volketal. (2011) measured changes in ecosystem C fluxes in response to the 10 and
50 kg N/ha/yr treatments. Although aboveground plant growth increased, there were no
significant changes in gross primary productivity or ecosystem respiration. Net
ecosystem production was unaffected by the lower rate of N addition, but declined with
the 50 kg N/ha/yr treatment, which suggests decreases in belowground C pools.
Bassin et al. (2013) and Yolk et al. (2014) reassessed this Swiss Alps grassland N
addition experiment after 7 years when the experiment concluded. Yolk et al. (2014)
observed that N additions had a strong positive effect on aboveground productivity,
ranging from a 31% stimulation caused by 5 kg N/ha/yr to a 60% increase caused by
50 kg N/ha/yr. Effects on belowground biomass were smaller, ranging from a 19%
increase for the 5 kg N treatment to a 24% increase for the 50 kg N treatment. Thus, N
additions increased the ratio of shoot-to-root biomass (Yolk et al.. 2014). This is
consistent with a 15N tracer study at the site, in which the N addition treatment increased
the proportion of the 15N tracer found in aboveground biomass (Bassin et al.. 2015).
However, the stimulation of plant N pools was relatively similar among aboveground
biomass (+31%), roots (+31%), necromass (+41%), and litter [+30%; (Bassin et al..
2015)1. Among functional groups, the longer term results were both similar and different
from the results observed after the initial 3 years (Bassin et al.. 2013). The effects ofN
addition on sedge biomass increased through time, with the highest rate of N addition
(50 kg N/ha/yr) increasing sedge biomass by 360%. In this same treatment, grass biomass
increased by 14%, forb biomass was unchanged, and legume biomass decreased by 43%.
Unsurprisingly, these changes resulted in more sedge-dominated plant communities. Two
aspects of this community change are notable: (1) changes were largely complete after
4-5 years, and (2) the lowest rate of N addition (5 kg N/ha/yr) was sufficient to cause
significant changes (Bassin et al.. 2013). This experiment also included an ozone
fumigation treatment, but there were no interactions between ozone and added N for
productivity or community composition (Yolk et al.. 2014; Bassin et al.. 2013; Bassin et
al.. 2007) and few N x ozone interactions for leaf physiological traits (Bassin et al..
2009). Two related studies on alpine grasslands have also been conducted in the Swiss
Alps by members of the same research group. At 10 sites, (Bassin et al.. 2012) found that
grassland aboveground productivity increased 30% with 2 years of NH4NO3 additions of
50 kg N/ha/yr. Blanke et al. (2012) planted seedlings of four local plant species into
intact grass/soil monoliths excavated from the same Swiss alpine grassland site as Bassin
et al. (2013). Bassin et al. (2009). and Bassin et al. (2007). Nitrogen addition
(50 kg N/ha/yr) treatments increased overall biomass of the unplanted vegetation in the
monoliths. Among the planted species, N additions increased biomass of the grass and
decreased biomass of the forb, but had no effect on the sedge or legume. Roots of the
6-63
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
grass, forb, and legume were colonized by arbuscular mycorrhizal fungi; N addition
increased mycorrhizal root colonization in the grass (Blanke et al.. 2012).
As in other environments, the plant growth response to N additions in alpine and Arctic
ecosystems varies among taxonomic groups. For example, in the Swiss subalpine
grassland experiment, Bassin et al. (2013) noted that within plant communities, sedge
(Cyperaceae) species tended to show especially strong positive responses to N additions.
As discussed in Appendix 5. Bowman et al. (2012) found N additions of 5, 10, and
30 kg N/ha/yr increased cover of the sedge Carex rupestris from 34 to 125% within a dry
meadow in Rocky Mountain National Park in Colorado. In the northern Caucasus
Mountains of Russia, Onipchenko etal. (2012) examined the response of four different
alpine plant communities to 90 kg N/ha/yr for 5 years. The N additions increased
vascular plant biomass in a lichen heath community, but not in grassland, meadow, or
snowbed communities. Within these communities, the response of individual plant
functional groups to N varied. The N additions increased sedge biomass in the lichen
heath, increased forbs and sedges in the meadow, increased grasses in the grassland, and
decreased legumes in the lichen heath and meadow. Shrubs were not significantly
affected by N additions. This variation among plant species was also observed in two
long-term N addition experiments in moist alpine meadows at the Niwot Ridge
Long-Term Ecological Research site in Colorado, wherein N additions (averaging either
229 or 288 kg N/ha/yr) for 7 or 11 years increased growth of the grass Deschampsia
cespitosa, but decreased growth of the perennial forb Geum rossii (Farrer et al.. 2015;
Farrer et al.. 2013). The decrease in Geum abundance was not necessarily due to
competitive exclusion: Geum decreased even in study plots where Deschampsia had been
removed (Farrer et al.. 2013). Instead, Geum showed effects of physiological stress with
N additions, with less C allocation to storage organs and lower concentrations of
nonstructural carbohydrates (Farrer et al.. 2013). In wet, moist, and dry alpine meadows
at Niwot Ridge, two decades of N additions (averaging -85 kg N/ha/yr) did not alter total
aboveground biomass, but significantly increased graminoid abundance in the moist and
wet meadows (Yuan et al.. 2016). In a meta-analysis of 51 studies conducted within the
Tibetan Plateau, (Fu and Shen. 2016) found that N additions increased total plant
biomass, aboveground biomass, belowground biomass, plant height, and foliar N
concentrations. However, changes in aboveground biomass varied by plant functional
type, with graminoids and sedges responding positively to added N and forbs and
legumes responding negatively.
In addition to the research on alpine tundra in Colorado, there have been several other
recent publications on the effect of N additions on plant productivity in tundra
ecosystems in North America. Additions of 70 kg N/ha/yr for 3 years to plots in a central
Utah alpine meadow increased aboveground plant production by 10-20%, but did not
6-64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
significantly impact soil respiration or soil C content (Gill. 2014). Given the large
contribution of root and mycorrhizal respiration to soil respiration, these results suggest a
change in C allocation toward aboveground growth. In shrub tundra in the Northwest
Territories of Canada, 8 years of N additions of 10 kg N/ha/yr had no effect on overall
aboveground plant biomass or the biomass of the dominant plant (Betula), but each of
these metrics was increased by N additions at a rate of 100 kg N/ha/yr (Zamin et al..
2014; Zamin and Grogan. 2012). Inflorescence production in Betula was unaffected at
both N addition rates. Elsewhere in the Northwest Territories, Churchland et al. (2010)
observed that a single year of N additions (100 kg N/ha/yr) increased foliar N
concentrations in shrub tundra plants. In an Alaskan tundra meadow, Kellev and Epstein
(2009) found that N additions (100 kg N/ha/yr for 3 years) increased foliar N
concentrations but had no effect on aboveground plant growth. Together, the results of
these studies in North America are consistent with the overall body N addition research in
tundra ecosystems (Table 6-6).
The influence of N additions on tundra plant biomass and productivity has also been
assessed in other areas of the world since 2008. In particular, numerous N addition
experiments have been conducted in alpine grasslands within the Tibetan Plateau region
of China (Fu and Shen. 2016). Although the effects of added N on plant growth clearly
vary by functional type (Fu and Shen. 2016). differences between functional groups are
not necessarily consistent between ecosystems. For instance, while (Fu and Shen. 2016)
found that added N increased sedge biomass and decreased forb biomass in their
meta-analysis of Tibetan plateau studies, 4 years of NH4NO3 additions at rates of 10, 20,
40, and 80 kg N/ha/yr caused significant increases in aboveground plant biomass and
cover of grasses and forbs and either decreases or no change in sedges (Zona et al..
2016). Species-specific changes in cover and foliar N concentration on the Tibetan
Plateau were observed even in studies using very high N addition rates (50, 150, or
300 kg N/ha/yr for 3 years) (Xiong etal.. 2016). Also on the Tibetan Plateau,
100 kg N/ha/yr over 2 years increased leaf litter N concentrations in three herbaceous
species, but also changed the biochemical composition of leaf litter by increasing the
concentration of cellulose and decreasing the concentration of lignin in a forb (Gentiana
straminea) and decreasing the concentration of hemicellulose in a sedge (Kobresia
humilis) (Zhu et al.. 2016b). While higher leaf litter concentrations of these structural
biochemicals were linked to slower decomposition overall within this experiment, N
additions did not significantly alter decomposition. As in other terrestrial ecosystems,
increases in plant biomass on the Tibetan Plateau appear to be more dependent on the
amount of N added (3.75 or 15 vs. 75 kg N/ha/yr) than the form of N added (NaNC>3,
[NH4]2S04, NH4NO3) (Song andYu. 2015).
6-65
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
In an Arctic shrub ecosystem in Greenland, Arens et al. (2008) studied the effects of N
additions at 5, 10, and 50 kg N/ha/yr over 3 years on ecosystem C fluxes and community
composition. Net ecosystem productivity (NPP), gross primary production (GPP), and
ecosystem respiration (ER) all increased in response to N additions, but these responses
were saturated beyond 10 kg N/ha/yr. The smallest rate of N addition caused both GPP
and ERto nearly double. However, soil respiration was not significantly affected by the
N addition treatments. Dominant vascular plant cover was highest in the 10 kg N/ha/yr
treatment. Nitrogen additions increased the cover of subdominant vascular plants from
3% in control plots to 8% in the highest N treatment. There were also other changes in
the cover of plant functional groups, including a 41% increase in deciduous shrubs in the
lowest N addition plots, a doubling of graminoid biomass in the moderate N treatment
relative to control, and large increases in forb biomass in the moderate and high N
addition treatment.
As in forests, lichens are an important component of plant communities and a significant
contributor to ecosystem function in Arctic and alpine tundra ecosystems. Since 2008,
there have been numerous observations of how lichens respond to added N in these
systems (Table 6-7). and these responses are broadly similar to those observed in forests.
Britton and Fisher (2010) observed increased tissue N concentrations in three of four
lichen species from Scottish heathlands over the course of a growing season with N
additions of 2.5 to 22.5 kg N/ha/yr. Hogan et al. (2010b) collected lichens from
27 heathland sites in the U.K. along a wet N deposition gradient (2-33 kg N/ha/yr) and
observed increased thalli N concentrations, decreased thalli P concentrations, and
increased phosphomonoesterase activity, an enzyme important for P acquisition. Hogan
et al. (2010a) found similar results in an N addition study. Armitage et al. (2014)
observed a decline in lichen cover in alpine heathland sites across an N deposition
gradient of 0.6 to 39.6 kg N/ha/yr in the North Atlantic region of Europe. Similarly, there
were decreases in lichen biomass or lichen cover with additional N in studies in Scotland
(Britton and Fisher. 2010). Sweden (Wardle et al.. 2013). Alaska (Kellev and Epstein.
2009). and Canada (Zamin et al.. 2014).
6-66
DRAFT: Do Not Cite or Quote

-------
Table 6-7 Alpine and Arctic tundra lichen growth and physiology responses to nitrogen added via atmospheric
deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Zamin et al. (2014)
Northwest
Territories
(Canada)
Shrub tundra (Vaccinium
vitisidaea, Rhododendron
subarcticum, Andromeda
polifolia)
Addition
10, 100
8
Lichen biomass
Low dose: not
significant
Hiqh dose:
decrease
Kellev and Epstein (2009)
Alaska
Tundra meadow (Dryas
integrifolia, Eriophorum
vaginatum, Carex spp.)
Addition
100
3
Lichen cover
Decrease
Wardleet al. (2013)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Empetrum
hermaphroditum,
Vaccinium spp.)
Addition
50
21
Lichen cover
Decrease
Armitaqe et al. (2014)
Europe (North
Atlantic)
Alpine heathlands
Ambient
0.6-39.6
n/a
Lichen cover
Decrease
Britton and Fisher (2010)
Scotland
Heathlands (Calluna
vulgaris)
Addition
2.5, 7.5,
12.5, 22.5
0.25
Lichen thalli mass
Three species: not
significant
Two species:
decrease
Hoqan et al. (2010a)
U.K.
Heathland lichen
(Cladonia portentosa)
Addition
8, 24, 56
4
Phosphomono-
esterase activity
Increase
Hoqan et al. (2010b)
U.K.
Heathland lichen
(Cladonia portentosa)
Ambient
2.32-32.8
n/a
Phosphomono-
esterase activity
Increase
June 2018
6-67
DRAFT: Do Not Cite or Quote

-------
Table 6 7 (Continued): Alpine and Arctic tundra lichen growth and physiology responses to nitrogen added via
atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Britton and Fisher (2010)
Scotland
Heathlands (Calluna
vulgaris)
Addition
2.5, 7.5,
12.5, 22.5
0.25
Thalli N %
Three species:
increase
One species: not
significant
Hoqan et al. (2010a)
U.K.
Heathland lichen
(Cladonia portentosa)
Addition
8, 24, 56
4
Thalli N %
Increase
Hoaan et al. (2010b)
U.K.
Heathland lichen
(Cladonia portentosa)
Ambient
2.32-32.8
n/a
Thalli N %
Increase
Hoqan et al. (2010b)
U.K.
Heathland lichen
(Cladonia portentosa)
Ambient
2.32-32.8
n/a
Thalli N %
Increase
Hoaan et al. (2010b)
U.K.
Heathland lichen
(Cladonia portentosa)
Ambient
2.32-32.8
n/a
Thalli P %
Decrease
ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; P = phosphorus; yr = year.
Notes: single studies are reported more than once if multiple endpoints or ecosystems were measured. References ordered by endpoint. Only statistically significant effects are listed
as increases or decreases.
June 2018
6-68
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
Because of the often sparse cover of vascular plants in tundra ecosystems, bryophytes are
often relatively larger components of plant communities than in other terrestrial systems.
As with lichens, bryophyte physiology and growth in tundra ecosystems can be impacted
by added N. Armitage et al. (2012) conducted an analysis of the tissue chemistry, growth,
and cover of the bryophyte Racomitrium lanuginosum in alpine heathlands at 36 sites in
Great Britain, Iceland, the Faroe Islands, and Norway. Estimated N deposition at these
sites ranged from ~1 kg N/ha/yr to nearly 40 kg N/ha/yr. Among environmental variables
that included temperature, precipitation, and grazing, N deposition was the largest
influence on tissue N and P concentrations. Notably, tissue N concentration increased
rapidly with N deposition rates up to 5 kg N/ha/yr, then increased at a slower rate with
higher levels of N deposition. Tissue N concentration was positively related to moss
shoot growth, but also positively related to shoot turnover and negatively related to moss
mat depth. These effects were nonlinear and especially pronounced at high tissue N
concentrations. Among other plant functional types at these sites, shrub growth decreased
with greater N deposition, graminoid growth increased, and forbs were not significantly
affected (Armitage et al.. 2014). In related work using a subset of these sites, Armitage et
al. (2011) conducted a 2-year transplant experiment that moved Racomitrium between
10 sites receiving higher N deposition (8.2-32.9 kg N/ha/yr) and a lower N site
(7.2 kg N/ha/yr). When the authors transplanted Racomitrium from elevated N deposition
sites to low N deposition sites and vice versa, tissue N concentration remained higher in
the plants from higher N sites, whereas moss moved from lower N to higher N sites
increased tissue N to nearly match levels in the moss natural to the site. Armitage et al.
(2011) also observed that moss at high N sites had higher shoot growth, whereas moss
transplanted to lower N sites increased in biomass due to decreasing tissue C:N and
slowing decomposition (Armitage et al.. 2011).
In high-latitude and high-elevation ecosystems, such as tundra, heathlands, and boreal
forests, plants in the Ericaceae family are often abundant. Notable members of the
Ericaceae family include blueberries and cranberries (both Vaccinium). Ericaceous plants
uniquely host ericoid mycorrhizae, which are distinct from ectomycorrhizae and
arbuscular mycorrhizae. Ericoid fungi are recognized for their ability to decompose large
organic molecules and absorb organic forms of soil N (Read et al.. 2004). This
specialization in organic N uptake could make these mycorrhizae more sensitive to
increased inorganic N availability. However, only a relatively small number of N addition
studies have quantified changes in mycorrhizal growth or physiology in ecosystems
supporting plants that host ericoid mycorrhizae (Table 6-8). Ishida and Nordin (2010)
added 12.5 or 50 kg N/ha/yr for 4 or 12 years to boreal forests in Sweden and observed
no change in the percentage of Vaccinium myrtillus roots colonized by ericoid
mycorrhizae, the number of ericoid fungal species per root tip, or ericoid community
composition, and observed mixed effects on ericoid species richness. Dean et al. (2014)
6-69
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
studied the response of root-associated fungi in alpine tundra in Colorado to 8 years of N
additions (29 kg N/ha/yr). Ericoid fungi abundance decreased and the overall community
composition of root-associated fungi changed, including a reduction in diversity and
richness. Further research is needed to understand the response of ericoid mycorrhizae to
added N.
Although Bouskill et al. (2014) observed that N additions increased fungal biomass in a
meta-analysis of high-latitude N addition experiments, other studies of fungal biomass in
tundra ecosystems have found negative (Xiong et al.. 2016; Farrer et al.. 2013; Wardle et
al.. 2013) or neutral responses (Sundqvist et al.. 2014). Bouskill et al. (2014) did not find
a significant effect of added N on total microbial biomass in their meta-analysis, with
individual studies reporting a negative effect (Farrer et al.. 2015). positive effect (Xiong
et al.. 2016; Zona et al.. 2016) or no effect (Zona et al.. 2016; Churchland et al.. 2010). In
a meta-analysis, Treseder (2008) found that N additions decreased bacterial biomass in
tundra ecosystems, but this analysis included only two studies. Of the four subsequent
studies of bacterial biomass, N additions had no effect in two, a positive effect in one,
and a negative effect in a fourth (Table 6-9). Notably, Bouskill et al. (2014) observed that
microbial biomass responses were positive under relatively low rates of N addition, but
became negative as N addition rates exceeded -50 kg N/ha/yr. Belowground respiration
showed a similar response, but lower threshold (-25 kg N/ha/yr) at which the response
switched from positive to negative.
To examine the influence of added N on organisms at higher trophic levels, Bishop et al.
(2010) studied the effects of N additions (78 kg N/ha/yr for 5 years) on plant and
arthropod communities in alpine meadows that were primary successional ecosystems on
recently formed volcanic substrates on Mount St. Helens in Washington. The N additions
increased foliar N concentration in the dominant graminoid, but decreased foliar N
concentration in the dominant legume. The N additions also increased both plant species
diversity and plant cover, although in the final year of the experiment, increased
browsing by small mammals negated the effect of the N additions. There was no effect on
total arthropod abundance (individuals per m2), but the number of Orthoptera
approximately doubled. Overall, total arthropod abundance was strongly positively
related to both plant diversity and plant cover, with arthropod abundance among
individual orders both positively and negative related to plant diversity and plant cover
6-70
DRAFT: Do Not Cite or Quote

-------
Table 6-8 Growth and biodiversity responses of ericoid mycorrhizal fungi to
nitrogen added in experimental treatments.




Nitrogen






Ambient
Addition


Effect of

Study

Deposition
Rate (kg
Duration

Additional
Reference
Location
Vegetation
or Addition
N/ha/yr)
(yr)
Endpoint
Nitrogen
Ishida and
Sweden
Vaccinium
Addition
12.5, 50
12
Ericoid
Not significant
Nordin (2010)

myrtillus roots in


(Picea),
community



Picea abies


4 (Pinus)
composition



forests and







Pinus sylvestris







forest





Dean et al.
Colorado
Alpine tundra
Addition
28.8
8
Ericoid fungal
Decrease
(2014)
(Niwot
(Geum rossii)



abundance


Ridge)






Ishida and
Sweden
Vaccinium
Addition
12.5, 50
12
Ericoid species
Not significant
Nordin (2010)

myrtillus roots in


(Picea),
per root tip



Picea abies


4 (Pinus)




forests and







Pinus sylvestris







forests





Ishida and
Sweden
Vaccinium
Addition
12.5, 50
12
Ericoid species
Pinus\
Nordin (2010)

myrtillus roots in


(Picea),
richness
increase


Picea abies


4 (Pinus)

Picea\ not


forests and




significant


Pinus sylvestris






forests





Ishida and
Sweden
Vaccinium
Addition
12.5, 50
12
Root
Not significant
Nordin (2010)

myrtillus roots in


(Picea),
colonization



Picea abies


4 (Pinus)
(%)



forests and







Pinus sylvestris







forests





Dean et al.
Colorado
Alpine tundra
Addition
28.8
8
Root-
Change
(2014)
(Niwot
(Geum rossii,



associated


Ridge)
Deschampsia



fungal



cespitosa)



community







composition

Dean et al.
Colorado
Alpine tundra
Addition
28.8
8
Root-
Deschampsia'.
(2014)
(Niwot
(Geum rossii,



associated
decrease

Ridge)
Deschampsia



fungal diversity
Geum\


cespitosa)



and richness
increase
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only
statistically significant effects are listed as increases or decreases.
June 2018
6-71
DRAFT: Do Not Cite or Quote

-------
Table 6-9 Alpine and Arctic tundra microbial biomass responses to experimental nitrogen additions.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Treseder (2008)
Tundra
Meta-analysis
Addition

Various
Bacterial
biomass
Decrease
Farreretal. (2013)
Colorado
(Niwot Ridge)
Moist alpine meadow
(Deschampsia cespitosa,
Geum rossii)
Addition
288
11
Bacterial
biomass
Decrease
Wardleet al. (2013)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Empetrum hermaphroditum,
Vaccinium spp.)
Addition
50
21
Bacterial
biomass
Not significant
Sundqvist et al. (2014)
Sweden
Tundra heath (Vaccinium
vitisidaea, Vaccinium
uiiginosum, Betuia nana)
Addition
100
3
Bacterial
biomass
Not significant
Farreretal. (2013)
Colorado
(Niwot Ridge)
Moist alpine meadow
(Deschampsia cespitosa,
Geum rossii)
Addition
288
11
Fungal biomass
Decrease
Wardleet al. (2013)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Empetrum hermaphroditum,
Vaccinium spp.)
Addition
50
21
Fungal biomass
Decrease
Bouskill et al. (2014)
North America
and Europe
Arctic and high-latitude meta-
analysis
Addition
Average: 72;
range: 1-100
Various
Fungal biomass
Increase
Sundqvist et al. (2014)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Anthoxanthum alpinum)
Addition
100
3
Fungal biomass
Not significant
June 2018
6-72
DRAFT: Do Not Cite or Quote

-------
Table 6 9 (Continued): Alpine and Arctic tundra microbial biomass responses to experimental nitrogen additions.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Bouskill etal. (2014)
North America
and Europe
Arctic and high-latitude meta-
analysis
Addition
Average: 72;
range: 1-100
Various
Microbial
biomass
Not significant
Farrer et al. (2015)
Colorado
(Niwot Ridge)
Moist alpine meadow
(Deschampsia cespitosa,
Geum rossii)
Addition
229
7
Microbial
biomass
Decrease
Churchland et al. (2010)
Northwest
Territories
(Canada)
Shrub tundra (Betula
glandulosa, Vaccinium
vitisidaea, Rhododendron
subarcticum)
Addition
100
1
Microbial
biomass C
Not significant
Fu and Shen (2017)
Tibetan Plateau
Meta-analysis
Addition
10-350
0-8
Microbial
biomass C
Not significant
C = carbon; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints or ecosystems were measured. References ordered by endpoint. Only statistically significant effects are listed
as increases or decreases.
June 2018
6-73
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
6.2.5.
Grasslands
The 2008 ISA contained limited information regarding the effects of N deposition on
productivity and related aspects of eutrophication in grasslands. That assessment included
the work of (Ncff et al.. 2002). who examined the long-term effects (+10 years) of N
deposition (10 kg N/ha/yr) in an alpine meadow, where N additions stimulated
aboveground NPP by 47%. Factors affecting productivity in grasslands are similar in
many ways to forests, with N availability limiting primary production. Grasslands
dominate largely in places where forests cannot grow because of low moisture
availability, high disturbance rates (such as from fire or grazing), and/or cold
temperatures and shorter growing seasons [such as at high latitudes and elevations;
(Chapin et al.. 2002; Knapp et al.. 1998)1. However, because of the commonality of N
limitation to primary production (Vitousek and Howarth. 1991). many of the processes
that occur when N is added to forests also occur when N is deposited onto grasslands.
These include increases in primary production and foliar N, increases in allocation to
aboveground biomass (increased shoot:root ratio), decreased levels of light reaching the
ground surface, elevated litter inputs, and changes in the soil biota (Aerts and Chapin.
1999; Knapp et al.. 1998). which are described further below.
Since the 2008 ISA, further advancements in our understanding of the effects of N
deposition on grasslands have occurred. A meta-analysis estimated that N additions
increased grassland aboveground NPP by 53% (LeBauer and Treseder. 2008). Increased
plant growth, NPP, or plant biomass in response to added N has been observed in prairies
in Wyoming, Kansas, and Minnesota (Henry et al.. 2015; Farrior et al.. 2013; Isbell et al..
2013a; Hillerislambers et al.. 2009; Johnson et al.. 2008). in temperate grasslands in
Ontario (Vankoughnctt and Henry. 2014; Hutchison and Henry. 2010) and Michigan
(Grinan and Robinson. 2013). in Mediterranean grasslands in California (Tulloss and
Cadenasso. 2016; Borer et al.. 2014; Vallano et al.. 2012). and steppe grasslands of China
(Li et al.. 2017a). Although the response to N enrichment is variable among species in
many terrestrial ecosystems, the comparatively rapid turnover rate of plant biomass in
grasslands means that wide differences in the response of individual species to N
deposition are frequently observed even in short-term studies (Tulloss and Cadenasso.
2016; Hautier et al.. 2014; Vankoughnett and Henry. 2014; Grman and Robinson. 2013;
Isbell et al.. 2013a; Bradford et al.. 2012; Vallano et al.. 2012; Skogen et al.. 2011). Three
new syntheses of grassland responses to N additions or N deposition have been published
since the 2008 ISA. Stevens et al. (2015) used data from a network of 42 grassland sites
on four continents and found that estimated rates of atmospheric N deposition were the
strongest predictor of aboveground NPP in these ecosystems. On average, aboveground
June 2018
6-74
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
NPP increased 3% for every 1 kg N/ha/yr increase in N deposition, and N deposition
explained 16% of the variance in productivity in these systems (Stevens et al.. 2015).
implying that much of the variation in NPP was not understood. Yue et al. (2016)
conducted a meta-analysis of N addition studies and found a 50.4% stimulation of
grassland aboveground NPP, similar to the estimate of LeBauer and Treseder (2008).
Aboveground plant C was stimulated by an average of 30.5% in the Yue et al. (2016)
analysis.
The belowground productivity responses to N appear to be more mixed. Yue et al. (2016)
found a 28% increase in belowground plant C; whereas Li et al. (2015) did not observe a
significant change in grassland fine root biomass. Belowground net primary productivity
in an experimental grassland declined with N additions when summed over a 9-year
period, along with the ratio of belowground-to-aboveground net primary productivity (Xu
et al.. 2017b). In a controlled pot experiment, Wang et al. (2017b) found that both fine
root production and turnover rates increased in a temperate grass species (Bothriochloa
ischaemum) with N additions. The net result was lower standing root biomass.
The differences in how individual plant species respond to added N may in part be due to
variation in functional traits, including the ability to change plant C allocation. In a
common garden experiment, Johnson et al. (2008) grew plants that were among the most
positively and negatively responsive to long-term N additions at the Cedar Creek and
Konza Prairie LTER sites. The plants that were most positively affected by N additions
allocated more biomass to shoots than roots and formed fewer associations with
mycorrhizal fungi. Likewise, Grman and Robinson (2013) also found that N additions
increased aboveground biomass and decreased the relative allocation of C to
mycorrhizae. Of the two grass species studied, the species that more strongly decreased
allocation to mycorrhizae as N availability increased also showed a greater aboveground
biomass response to added N. Bradford et al. (2012) documented a somewhat different
functional response, finding that increasing rates of N additions favored the species that
showed greater leaf trait plasticity in response to the N additions. Earlier studies
suggested leaf plasticity was an important determinant of plant growth responses to added
N, and that the ability of a species to upregulate production if nutrients were abundant
was an advantage (Knops and Reinhart. 2000). Hillerislambers et al. (2009) observed that
N amendments (40 kg N/ha/yr) had similar effects on grasses within the same functional
groups: C3 grasses tended to increase aboveground biomass and decrease seed
production, while C4 grasses increased seed production and decreased aboveground
biomass. In a survey of 44 species in 153 acidic grassland sites across northern and
western Europe, Pannek etal. (2015) found that species with high relative growth rates
tended to show beneficial responses to N deposition.
6-75
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
In multiyear studies, responses to N enrichment often varied through time. For instance,
over a 7-year period of N addition, Henry et al. (2015) found that the stimulation of
aboveground biomass ranged approximately 10 to nearly 100%. Working at a long-term
biodiversity and N deposition experiment at Cedar Creek in Minnesota that used seven
different N addition rates from 10 to 270 kg N/ha/yr, Isbell et al. (2013a) initially
observed that N additions increased grassland productivity, with higher rates of N
addition causing greater increases in productivity. Although N additions continued to
stimulate productivity, productivity decreased in plots that lost the most species over
time. In a large synthesis, Hautier et al. (2014) used data from studies of 41 experimental
grassland communities on 5 continents and found grassland communities had greater
annual variability in plant productivity (less stability) as the N input rate increased. Under
ambient conditions, annual variability in plant productivity is limited by asynchronous
productivity among different species, wherein slower growth by some species is balanced
by more rapid growth of neighboring species. In these unmanipulated systems, the
stability of plant productivity is positively associated with plant diversity (Hautier et al..
2014; Tilman et al.. 2001; Tilman et al.. 1997). With added N, the increase in annual
variability in productivity could not be linked to a loss of species richness, but was
instead caused by a decrease in the asynchrony of productivity among individual plant
species. In a more recent experiment in China, Zhang et al. (2016a) observed that N
additions (10-500 kg N/ha/yr for 6 years) created a series of changes: decreased plant
species richness, decreased asynchrony among plant species, increased aboveground
productivity, and decreased stability (more annual variability) in aboveground
productivity.
Nitrogen enrichment can also cause variation in plant chemistry, particularly increases in
tissue N concentrations (Lu et al.. 2017; Reich et al.. 2003). (Bradford et al.. 2012)
conducted a greenhouse experiment with several grass and forb species and observed
increases in shoot N concentrations, but root C:N and root mass N were not consistently
affected by N additions. In a greenhouse experiment conducted by (Jamicson et al..
2012). N addition had no effect on the concentration of a defensive compound in shoots,
whereas concentrations decreased in flowers by -35%. As in other terrestrial ecosystems,
these increases in shoot N concentrations can lead to an increase in leaf-level
photosynthesis if other resources are not limiting [e.g., (Reich et al.. 2003; Lee et al..
2001)1. However, such increases in photosynthesis vary among species and can vary
temporally (Reich et al.. 2003; Lee et al.. 2001). Within a U.K. grassland exposed to N
additions of 35 or 140 kg N/ha/yr for 11 years, Arroniz-Crespo et al. (2008) observed
decreases in cover of two bryophyte species as well as a decrease in chlorophyll
fluorescence in both species (despite an increase in chlorophyll content in one species).
Both bryophyte species exhibited increased activity of an enzyme involved in P
acquisition (phosphomonoesterase) and decreased nitrate reductase enzyme activity, but
6-76
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
only one of the two bryophyte species exhibited increased tissue N concentrations and
N:P ratios.
Several experiments in Ontario, Canada have examined the response of old field
grasslands to the interaction between N additions and climate change in the form of
altered snowpack, soil freezing, or warming. Vankoughnett and Henry (2014) found that
N additions of 20 or 60 kg N/ha/yr increased aboveground production, but this increase
varied by species. In the first year of the experiment, plant production showed a stronger
response to N in plots where snow had been experimentally removed, but this interaction
did not occur in subsequent years. Although the extent of the effect varied by year, Henry
et al. (2015) found that 7 years of N additions at a rate of 60 kg N/ha/yr increased
aboveground biomass and showed no interaction with experimental warming.
Effects of water availability were also observed in grassland studies by Friedrich et al.
(2012). Jamieson et al. (2013). and Farrior et al. (2013). In a greenhouse experiment with
a grass native to German grasslands, Friedrich et al. (2012) found that N additions of
48 kg N/ha/yr increased aboveground biomass by 500%. However, plants receiving N
were more likely to suffer dieback when N additions were combined with drought.
Nitrogen enrichment in a Wyoming prairie increased the production of a defensive
chemical compound in an invasive plant by 37% under ambient conditions, but decreased
the production of this compound by 25% when water availability was reduced (Jamieson
et al.. 2013). At Cedar Creek in Minnesota, Farrior et al. (2013) found complex responses
to changes in N and water availability: N enrichment increased shoot biomass regardless
of water availability, fine root biomass declined when availability of both water and N
was high, and higher water availability increased fine root biomass only at low rates of N
addition.
There has been less research on the effects on N enrichment on belowground C cycling in
grasslands than in forests (Yue et al.. 2016; Li et al.. 2015; Liu and Greaver. 2010). In a
meta-analysis, Liu and Greaver (2010) found that N additions to grasslands increased
aboveground litter inputs, but did not significantly affect soil respiration, microbial
respiration, or soil C. In part, the lack of an overall effect on these processes could be due
to the combination of high variability among grasslands and the small sample sizes for
the meta-analysis, as only approximately six studies were available for the Liu and
Greaver (2010) analysis. However, a more recent meta-analysis of N addition
experiments by Yue et al. (2016) that had larger sample sizes found somewhat similar
results: belowground plant C was stimulated by 24.2%, but there were no significant
effects on soil respiration, microbial respiration, litter decomposition rates, or soil organic
C. Meta-analyses of changes in microbial biomass have produced inconsistent results:
Yue etal. (2016) did not find a significant effect of added N, Liu and Greaver (2010)
6-77
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
found a negative effect, and Geisseler et al. (2016) observed a negative effect only with at
least 5 years of N additions.
The conflicting meta-analytic results for changes in grassland soil microbial biomass
could be caused by differences in analyses or data sets, or could reflect ecological
complexities that require more nuanced understanding, such as the influence of P
availability, varying responses among microbial taxa, or measurement techniques
(Johnson et al.. 2008). For instance, at Jasper Ridge in the San Francisco Bay area, Liang
et al. (2015) did not find any significant changes in fungal lipid biomass caused by N
additions (70 kg N/ha/yr), yet did observe changes in the abundance of individual amino
sugars that indicated decreases in total microbial biomass and fungal biomass, and
increases in bacterial biomass. In measuring the abundance of nitrite-oxidizing bacteria at
Jasper Ridge, Le Roux et al. (2016) observed that N additions increased the abundance of
Nitrobacter, but did not affect Nitrospira. In the over 150-year Park Grass Experiment,
Zhalnina et al. (2015) observed that NaNC>3 additions had no effect on bacterial and
archaeal biomass as assessed by 16S ribosomal RNA abundance, but (NFL^SC^
additions decreased bacterial and archaeal biomass. Li et al. (2015) found no effect of N
additions on fine root biomass or mycorrhizal colonization of root tips in a meta-analysis
of grassland experiments, but sample sizes were small (four biomass studies, nine
mycorrhizal studies).
A number of studies have examined the effects of N additions on mycorrhizal abundance
in grasslands (Table 6-3). Grassland plants predominantly host arbuscular mycorrhizal
fungal associations. Like arbuscular mycorrhizal responses observed in forests [e.g., (van
Diepen et al.. 2010)1. arbuscular mycorrhizal responses to added N in grasslands have
been inconsistent, with a similar number of studies showing no response (Liang et al..
2015; Chen et al.. 2014; Mandvam and Jumpponen. 2008) versus showing a negative
effect on colonization or growth [e.g., (Chen et al.. 2014; Johnson et al.. 2008; Van Per
Heiiden et al.. 2008)1.
Most studies of grassland soil microbial responses to N deposition (Table 6-10) have
focused on free-living soil heterotrophs or mycorrhizal fungi, but Weese et al. (2015)
examined the response of N fixing rhizobial bacteria in clover (Trifolium) as part of a
22-year chronic N addition (123 kg N/ha/yr as NH4NO3) experiment in a Michigan
grassland. Clover inoculated with rhizobial bacteria from soil or bacterial strains that
were collected from the long-term N addition experiment exhibited lower chlorophyll
content and decreased plant biomass, including fewer leaves and stolons. These patterns
were consistent across all three clover species used in the experiment. Notably, these
changes did not result from a shift in the community composition of the rhizobial
bacteria, but instead resulted from the evolution of new strains of the bacteria Rhizobium
6-78
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
leguminosarum that were less symbiotic and less cooperative in the exchange of C for N.
A similar change toward less mutualistic interactions in response to N additions have
been observed for mycorrhizae in grasslands (Johnson. 1993). It is unclear whether these
evolutionary changes are reversible, and if so, how long this reversion would take (Weese
et al.. 2015).
Zehnder and Hunter (2008) performed a plant density manipulation experiment to
examine the effect of N deposition on the interaction between a host plant from the
southeastern U.S., Asclepias tuberosa, and its herbivore, Aphis nerii. Added N increased
aphid population growth, plant foliar N concentrations, and plant biomass. In southern
California coastal grasslands, Borer et al. (2014) observed that N additions (40 or
100 kg N/ha/yr as Ca|NO,|2) increased plant biomass, but only in the absence of
herbivores (pocket gophers, Thomomys bottae).
6-79
DRAFT: Do Not Cite or Quote

-------
Table 6-10 Grassland microbial biomass responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Eisenhauer et al. (2012)
Minnesota (Cedar
Creek)
Prairie C3 and C4
grasses, forbs,
legumes
Addition
40
14
Amoeba and
flagellate
abundance
Not significant
Kastl et al. (2015)
Germany (green-
house)
Temperate grasses
(Dactylis glomerata,
Festuca rubra)
Addition
50, 100,
200
0.12
Archaeal and
bacterial amoA
(NHs
monooxygenase)
gene abundance
Achaea: not
significant
Bacteria: increase
Wei et al. (2013)
China
Steppe grassland
Addition
5.6, 11.2,
22.4, 39.2,
56
4
Bacterial biomass
Lowest dose: not
significant
Other doses:
decrease
Liana et al. (2015)
California (northern)
Annual grassland
(Avena barbata, A.
fatua)
Addition
70
9
Bacterial biomass
Not significant
Wei et al. (2013)
China
Steppe grassland
Addition
5.6, 11.2,
22.4, 39.2,
56
4
Fungal biomass
Three lowest
doses: not
significant
Two hiahest doses:
decrease
Liana et al. (2015)
California (northern)
Annual grassland
(Avena barbata, A.
fatua)
Addition
70
9
Fungal biomass
Not significant
June 2018
6-80
DRAFT: Do Not Cite or Quote

-------
Table 6 10 (Continued): Grassland microbial biomass responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Ramirez et al. (2010b)
Minnesota (Cedar
Creek)
Temperate
grasslands
Addition
30, 60,
100, 160,
280, 500,
800
27
Microbial biomass
Not significant
Eisenhauer et al. (2013)
Minnesota (Cedar
Creek)
Prairie
Addition
40
14
Microbial biomass
Not significant
Wei et al. (2013)
China
Steppe grassland
Addition
5.6, 11.2,
22.4, 39.2,
56
4
Microbial biomass
Lowest dose: not
significant
Other doses:
decrease
Liana et al. (2015)
California (northern)
Annual grassland
(Avena barbata, A.
fatua)
Addition
70
9
Microbial biomass
Not significant
Li et al. (2017a)
China (Inner
Mongolia)
Steppe grassland
Addition
50, 100,
150
8
Microbial biomass
C
Decrease
Liana et al. (2015)
California (northern)
Annual grassland
(Avena barbata, A.
fatua)
Addition
70
9
Saprotrophic
fungal biomass
Not significant
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints or ecosystems were measured. References ordered by endpoint. Only statistically significant effects are listed
as increases or decreases.
June 2018
6-81
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
6.2.6.
Arid and Semiarid Ecosystems
Large portions of the western U.S. are covered by arid and semiarid ecosystems. While
many of these ecosystems receive little anthropogenic N deposition compared to the
eastern U.S., some areas downwind of agricultural and metropolitan centers receive high
levels of atmospheric N deposition (Fenn et al.. 2003b'). In addition, the unique nutrient
cycling processes in these systems can intensify the influence of N deposition on
ecosystem processes lYHomvak et al.. 2014); see Appendix 41.
The lack of moisture in arid and semiarid climates changes the nature of how N
deposition impacts biological communities in these ecosystems compared to more mesic
ecosystems such as tallgrass prairies and forests. First, the strong moisture constraint on
productivity limits the influence of increased N availability on plant communities
[e.g., (Rao and Allen. 2010)1. with shifts in community composition less likely to occur
without sufficient moisture [e.g., (Concilio and Loik. 2013)1. Second, the lack of strong
biological demand for N during periods of moisture limitation, combined with infrequent
soil leaching and runoff events, leads to the accumulation of inorganic N in surface soils,
intensifying the effects of N deposition on soil N concentrations when moisture is
available [e.g., (Homvak et al.. 2014)1. Broadly, soil base saturation and pH increase as
the climate becomes more arid (Schlesinger. 2005). buffering dryland ecosystems from
the acidification-driven losses in biodiversity that are observed along N deposition
gradients [e.g., (Simkin et al.. 2016)1. Finally, the patchy vegetation that often develops
in deserts tends to create isolated islands of fertility, wherein nutrients (including N) are
concentrated beneath shrub canopies (Titus etal.. 2002; Schlesinger and Pilmanis. 1998).
Under these circumstances, anthropogenic N deposition can increase N availability in the
interspaces between shrubs, allowing the growth of annual grasses and forbs, which can
grow and reproduce during the brief seasonal periods with adequate moisture availability
(Brooks. 2003). In particular, there have been numerous observations from semiarid and
arid ecosystems that N deposition can increase the productivity and dominance of
invasive grasses [e.g., (Brooks. 2003; Fenn et al.. 2003a; Padgett and Allen. 1999)1. This
phenomenon provides a more continuous fuel bed for wildfires, increasing the potential
for fire (Rao et al.. 2010) and shifting plant community composition away from species
that are not fire adapted (Padgett and Allen. 1999; Allen et al.. 1998).
There was already a large amount of information available on how N deposition impacted
arid and semiarid ecosystems at the time of the 2008 ISA, especially in areas around
southern California downwind of Los Angeles, where dry N deposition can be
>30 kg N/ha/yr (Bvtnerowicz and Fenn. 1996). In particular, the 2008 ISA described
results from several N deposition studies in this region that showed increased biomass of
June 2018
6-82
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
invasive species, high mortality of native shrubs in heavily polluted areas, and increased
fire risk (Padgett and Allen. 1999; Allen et al.. 1998). These experiments were primarily
conducted in coastal sage scrub (CSS) ecosystems or in the Mojave Desert.
The amount of land area covered by CSS in southern California has greatly declined in
the past 60 years due to changes in land use, grazing, and increased wildfire frequency
(Allen etal.. 1998). In addition, native CSS vegetation has been replaced in many areas
by annual grasses from the Mediterranean region (Padgett et al.. 1999; Padgett and Allen.
1999; Allen et al.. 1998). In the mid-1990s, N deposition had been implicated as a
contributor to this ecosystem alteration because it decreased the growth of native shrubs
and increased the growth of invasive grasses, with the secondary effect of increasing fire
frequency (Padgett et al.. 1999; Padgett and Allen. 1999; Allen etal.. 1998). There was
also evidence that the N assisted conversion from CSS to invasive annual grasses had
altered the hydrology of these ecosystems (Wood et al.. 2006). The loss of native
vegetation in the CSS and chaparral ecosystems in southern California is particularly
notable because these ecosystems are unique within the U.S., are limited in their spatial
extent, and are important global and national hotspots for biodiversity.
At the time of the 2008 ISA, similar effects of N deposition on plant communities had
also been observed farther east in the Mojave Desert. Brooks (2003) documented that
NH4NO3 additions (32 kg N/ha/yr) increased the biomass of invasive annual grasses and
forbs by more than 50%. The increased growth of the invasive plants suppressed the
growth of native annual plants. In particular, N additions stimulated the growth of
Bromus madritensis beneath the dominant native shrub Larrea tridentata, while the
invasive grasses in the genus Schismus and the invasive forb Erodium cicutarium had
enhanced growth in the interspaces between shrubs. This spatial nature of these effects is
important because as in the CSS, the spatially continuous fuel beds created by high grass
biomass have been associated with increased fire frequency in the Mojave Desert (Brooks
et al.. 2004; Brooks and Esque. 2002; Brooks. 1999). This effect is stronger at higher
elevation, likely because the higher precipitation stimulates grass productivity. Fire was
relatively rare in the Mojave Desert until the 1980s (Svphard et al.. 2017). but now fire
occurs frequently in areas that have experienced invasion of exotic grasses (Brooks.
1999). The comparatively unresponsive nature of native Mojave vegetation to N
deposition was also observed elsewhere. For instance, 3 years of N additions
[40 kg N/ha/yr as Ca(NO,)2| had almost no impact on leaf-level photosynthetic or
hydraulic performance in Larrea tridentata growing north of Las Vegas, NV (Barker et
al.. 2006).
At the time of the 2008 ISA, N deposition effects on belowground processes in arid and
semiarid lands were not well understood. In a Chihuahuan Desert grassland in New
6-83
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Mexico, 10 years of N additions (10 kg N/ha/yr) increased the activity of extracellular
glucosidase enzymes important in the breakdown of carbohydrates, but decreased the
activity of extracellular aminopeptidase enzymes important in the decomposition of
proteins and peptides (Stursova et al.. 2006). In contrast, there were no differences in C
mineralization (decomposition) in soils collected at a similar southern California gradient
of CSS sites [4 to 23 kg N/ha/yr; (Vourlitis and Zorba. 2007)1. Along a CSS N deposition
gradient in the Los Angeles area, arbuscular mycorrhizal fungi associated with the
dominant sagebrush shrub declined with increasing N pollution and also in response to N
additions (60 kg N/ha/yr; 4-5 years) at a relatively unpolluted site (Sigi'icnza et al.. 2006).
Evidence is conflicting about whether these changes in mycorrhizal colonization
influenced the plant growth response to N deposition. An earlier study suggested that
differences in mycorrhizal colonization were not responsible for the effects of N
deposition on CSS plant growth in this region (Yoshida and Allen. 2001). Sigi'icnza et al.
(2006) found that sagebrush inoculated with arbuscular mycorrhizae from a high N
deposition site grew more slowly than those with inoculum from a low N deposition site.
Grasses in semiarid ecosystems in New Mexico and Colorado grew slower when
inoculated with mycorrhizae from sites that had received N additions than from sites
without N additions (Corkidi et al.. 2002).
Research published since 2008 has continued to concentrate on CSS and Mojave Desert
ecosystems in the southern California, but new studies in southern California chaparral
ecosystems, the Sonoran Desert of Arizona, and elsewhere have also been published.
Most studies investigating the effects of N deposition in arid and semiarid environments
published since the 2008 ISA have further emphasized that the impacts of increased N are
heavily dependent on water availability (Vallicrc and Allen. 2016a; Rao et al.. 2015;
Homvak et al.. 2014; Newingham et al.. 2012; Ochoa-Hueso and Manrique. 2010; Rao
and Allen. 2010). The high dependence on precipitation is consistent with ecosystem
models for C cycling in desert systems (Shen et al.. 2008).
For instance, Rao and Allen (2010) investigated how productivity in Joshua Tree
National Park annuals is altered by increasing N supply under a range of water
availabilities (Table 6-11). Using a 5-year field N addition experiment and N additions of
up to 30 kg N/ha/yr, the authors observed the greatest production of invasive grasses and
native forbs under the highest soil N and highest watering regime. Similarly, Newingham
et al. (2012) found that N additions of 10 or 40 kg N/ha/yr as Ca(NO,)2 in the Mojave
Desert increased branch production in creosote (Larrea tridentata) only during a wet
year, whereas the addition of water and N significantly increased the amount of rodent
herbivory.
6-84
DRAFT: Do Not Cite or Quote

-------
Table 6-11 Arid and semiarid ecosystem plant productivity and physiology responses to nitrogen added in
experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of Additional
Nitrogen
Rao and Allen (2010)
Garden
(California,
Mojave)
Invasive grass (Bromus
madritensis), native forb
(.Amsinckia tessellata)
Addition
5, 30
1
Aboveground
plant biomass
Forb: increase
Grass: increase
Wanqetal. (2015b)
China (north,
Jilin)
Arid grassland (Leymus
chin en sis)
Addition
100
4
Ecosystem
respiration
Increase
Zhanq et al. (2015a)
Beijing, China
Shrubland (Vitex negundo)
Addition
20, 50, 100
1
Foliar N %
Low dose: not
significant
Mid dose: increase in
one of four species
Hiqh dose: increase in
3/4 species
Zhanq et al. (2015a)
Beijing, China
Shrubland (Spirea
trilobata)
Addition
20, 50, 100
1
Foliar N %
Not significant
Rao and Allen (2010)
California
(Mojave)
Larrea tridentata or
Juniperus californica, Pinus
monophylla
Addition
2, 5, 30
5
Grass and
forb biomass
Increase
Wanqetal. (2015b)
China (north,
Jilin)
Arid grassland (Leymus
chin en sis)
Addition
100
4
Gross
ecosystem
production
Increase
Pasquini and Vourlitis
(2010)
California
(southern;
three sites)
Chaparral (Adenostoma
fascicuiatum, Ceanothus
spp.)
Ambient
8.1, 11.9,
18.4
n/a
Growth rate
per shrub
Increase
June 2018
6-85
DRAFT: Do Not Cite or Quote

-------
Table 6 11 (Continued): Arid and semiarid ecosystem plant productivity and physiology responses to nitrogen
added in experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of Additional
Nitrogen
Halletal. (2011)
Phoenix, AZ
Creosote and bursage
shrublands (Larrea
tridentata, Ambrosia spp.)
Addition
60
5
Herbaceous
annual plant
production
Increase
Allen et al. (2009)
California
(Joshua Tree
NP; four sites)
Creosote bush scrub
(Larrea tridentata)', pinyon-
juniper woodland (Pinus
monophylla, Juniperus
caiifornica)
Addition
5, 30
2
Invasive
grass
biomass
Low dose: not
significant
Hiah dose: increase at
3/4 sites
Newinaham et al. (2012)
Nevada
(Mojave)
Creosote bush (Larrea
tridentata)
Addition
10, 40
6
Larrea
branch
elongation
Not significant
Newinqham et al. (2012)
Nevada
(Mojave)
Creosote bush (Larrea
tridentata)
Addition
10, 40
6
Larrea
branch
number
Increase
Newinaham et al. (2012)
Nevada
(Mojave)
Creosote bush (Larrea
tridentata)
Addition
10, 40
6
Larrea leaf
production
Not significant
Newinaham et al. (2012)
Nevada
(Mojave)
Creosote bush (Larrea
tridentata)
Addition
10, 40
6
Larrea seed
production
Not significant
Zhana et al. (2015a)
China (north,
Songnen)
Shrubland (Vitexnegundo)
Addition
20, 50, 100
1
Leaf litter N
%
Low dose: not
significant
Mid dose: increase in
one of four species
Hiah dose: increase
Zhana et al. (2015a)
China (north,
Songnen)
Shrubland (Spirea
triiobata)
Addition
20, 50, 100
1
Leaf litter N
%
Not significant
June 2018
6-86
DRAFT: Do Not Cite or Quote

-------
Table 6 11 (Continued): Arid and semiarid ecosystem plant productivity and physiology responses to nitrogen
added in experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of Additional
Nitrogen
Allen et al. (2009)
California
(Joshua tree
NP; four sites)
Creosote bush scrub
(Larrea tridentata)', pinyon-
juniper woodland (Pinus
monophylla, Juniperus
californica)
Addition
5, 30
2
Native and
exotic plant
cover
Low dose: not
significant
Hiah dose: not
significant at three sites,
increased native cover
at one site
Wanqetal. (2015b)
China (north,
Jilin)
Arid grassland (Leymus
chin en sis)
Addition
100
4
Net
ecosystem
exchange
Increase
Belnap et al. (2008)
Utah (Canyon-
lands NP)
Biological soil crusts
Addition
Not stated
1
Photosyn-
thetic
pigments
Decrease
Sun et al. (2014)
China (north,
Songnen)
Shrubland (Leymus
chinensis, Artemisia
scoparia)
Addition
23, 46, 69,
92
3
Plant
aboveground
biomass
Low dose: not
significant
Other doses: increase
Wanq et al. (2015b)
China (north,
Jilin)
Arid grassland (Leymus
chinensis)
Addition
100
4
Plant
aboveground
biomass
Increase
Zhana et al. (2015b)
China (north,
Songnen)
Alkaline grassland
(Leymus chinensis,
Kaiimeris integrifoiia)
Addition
100
4
Plant
aboveground
biomass
Increase
Collins et al. (2017)
New Mexico
Grassland (Bouteioua
eriopoda, Bouteioua
gracilis)
Addition
20
1 to 7
Plant
aboveground
biomass for
grass and
forbs
Increase in grass
biomass in 1 yr (the
year following a fire);
not significant in the
other years or for forb
biomass
June 2018
6-87
DRAFT: Do Not Cite or Quote

-------
Table 6 11 (Continued): Arid and semiarid ecosystem plant productivity and physiology responses to nitrogen
added in experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of Additional
Nitrogen
Zhana et al. (2015b)
China (north,
Songnen)
Alkaline grassland
(Leymus chinensis,
Kalimeris integrifolia)
Addition
100
4
Plant
belowground
biomass
Increase
Rao et al. (2009)
California
(Joshua Tree
NP)
Creosote bush (Larrea
tridentata) or pinyon-
juniper woodland (Pinus
monophylla, Juniperus
californica)
Ambient
2.7-14.4
n/a
Plant
biomass
Not significant
Ochoa-Hueso and Stevens
(2015)
Spain
Shrubland (Quercus
coccifera, Rosmarinus
officinalis, Lithodara
fruticosa)
Addition
10, 20, 50
3
Plant
biomass
Not significant
Rao et al. (2015)
California
(Mojave)
Desert wash, desert scrub,
desert succulent cover
types
Ambient
0.4-15.3
n/a
Plant
biomass
Not significant;
precipitation was main
driver of biomass
McHuqh et al. (2017)
Utah
Semiarid grassland
Addition
2, 5, 8
2
Plant cover
Not significant
Vourlitis (2017)
California
(southern)
Shrubland (coastal sage
scrub)
Addition
50
13
Plant cover
(total)
Not significant for most
of the 13 yr; increase in
the 7th and 8th yr of
experiment
Requs et al. (2017)
California
Desert, shrubland
Addition
1.8-8.7+
8 weeks
Plant growth
due to
rhizobia
nodules on
native
legume
(.Acmispon
strigosus)
Decrease
June 2018
6-88
DRAFT: Do Not Cite or Quote

-------
Table 6 11 (Continued): Arid and semiarid ecosystem plant productivity and physiology responses to nitrogen
added in experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration (yr)
Endpoint
Effect of Additional
Nitrogen
Halletal. (2011)
Phoenix, AZ
Creosote and bursage
shrublands (Larrea
tridentata, Ambrosia spp.)
Addition
60
5
Shrub
biomass
production
Not significant
Pasauini and Vourlitis (2010)
California
(southern;
three sites)
Chaparral (Adenostoma
fasciculatum, Ceanothus
spp.)
Ambient
8.1, 11.9,
18.4
n/a
Shrub
density
(plants/ha)
Decrease
Sinsabauqh et al. (2015)
Nevada
(Mojave)
Creosote and bursage
shrublands (Larrea
tridentata, Ambrosia
dumosa)
Addition
7, 15
1
Shrub foliar
element %
(including N,
P)
Not significant
Halletal. (2011)
Phoenix, AZ
Creosote and bursage
shrublands (Larrea
tridentata, Ambrosia spp.)
Addition
60
5
Shrub foliar
N %
Larrea and herbs:
increase
Ambrosia: not
significant
Wissinqer et al. (2014)
California
(Mojave)
Creosote and bursage
shrublands (Larrea
tridentata, Ambrosia
dumosa)
Ambient
2-12
n/a
Shrub foliar
N %
Not significant
Wissinqer et al. (2014)
California
(Mojave)
Creosote and bursage
shrublands (Larrea
tridentata, Ambrosia
dumosa)
Ambient
2-12
n/a
Shrub seed
production
Increase
ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; NP = national park; P = phosphorus; yr = year.
Notes: single studies are reported more than once if multiple endpoints or ecosystems were measured. References ordered by endpoint. Only statistically significant effects are listed
as increases or decreases.
June 2018
6-89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
A similar interaction between N and water was found in the Sonoran Desert. Herbaceous
annual biomass increased with added N (60 kg N/ha/yr) during above-normal rainfall
seasons, but not in low precipitation years (Hall et al.. 2011). In contrast, no increases in
productivity were observed in the dominant shrub Larrea tridentata, even during
above-normal rainfall (Hall et al.. 2011). Ecosystem modeling for this same study system,
however, conducted by Shen et al. (2008) suggested observed rates of N deposition in the
Phoenix region would increase Larrea productivity by more than an order of magnitude
in wet years. Mosses in this region also appear to be influenced by N deposition, with
lower moss abundance and higher tissue N concentrations at urban study sites that
receive higher N deposition (Ball and Guevara. 2015).
As in other terrestrial ecosystems (see Appendix 6.2.2. Appendix 6.2.3.1). increases in
foliar N concentrations in ecosystems receiving chronic N additions are not necessarily
associated with changes in leaf physiological function in semiarid plants. With the
exception of the N fixing Ceonothus, shrubs in a chaparral ecosystem and a CSS
ecosystem in southern California increased foliar N concentrations in response to
long-term N additions (50 kg N/ha/yr for 10 years), but none of the four shrub species
(Artemisia californica, Salvia mellifera, Adenostoma fasciculatum, Ceanothus greggii)
increased leaf-level photosynthesis (Pivovaroff et al.. 2016). Several species in this study
had higher dry season predawn water potential, indicating relative improvements in water
availability. However, wet season water potential (predawn and midday) were
unresponsive to N additions and only Artemisia exhibited changes in plant hydraulic
function (increased stomatal conductance and sapwood conductivity and decreased wood
density) that would indicate the shrubs were benefitting from greater water availability.
In contrast, Adenostoma wood xylem became more prone to cavitation (Pivovaroff et al..
2016). The limited aboveground physiological responses of the dominant shrubs in these
ecosystems is consistent with the idea that other resources (e.g., water) are often more
limiting to plant productivity than N in arid and semiarid ecosystems.
Other studies document the overall effect ofN deposition on vegetation communities,
including the occurrence of fire. At Joshua Tree National Park in the Mojave desert of
California, non-native grass biomass increased significantly at three of the four study
sites receiving 30 kg N/ha/yr for 2 years, but saw no change with 5 kg N/ha/yr of added
N (Allen et al.. 2009). Native species showed no clear response to N addition, but these
plants made up a small fraction of community biomass. Rao et al. (2015) tried to model
the effects of N deposition on the biomass of invasive annual grasses and the occurrence
of wildfire in the Mojave Desert of California, but were unable to effectively disentangle
the effects of N deposition on fire from factors such as precipitation, distance from roads,
and firefighting resources. However, in an earlier modeling study of Mojave Desert
vegetation, Rao et al. (2010) did conclude that increases in precipitation and N deposition
6-90
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
raise fire risks by stimulating greater vegetation biomass, and thus fuel loads. In a study
covering California deserts stretching from the Mexican border into the Great Basin, the
stimulatory effect of N deposition on the growth of annual grasses was one of the largest
contributors to the occurrence of large fires (>20 ha) across four semiarid ecoregions,
including the Mojave Desert (Svphard et al.. 2017). Additionally, Wissinger et al. (2014)
observed higher seed production in areas of the Mojave with higher atmospheric N
deposition (up to 16 kg N/ha/yr).
In arid ecosystems with low vascular plant cover, soil crusts with photoautotrophs such as
cyanobacteria, green algae, bryophytes, and lichens can be important ecosystem
components and large contributors to biogeochemical cycles. In the Mojave Desert, Stark
et al. (2011) investigated the effects of 10 or 40 kg N/ha/yr N addition on the dominant
biological soil-crust moss (Syntrichia caninervis) for a period of 5 years. Interestingly,
the lower rate of N application had negative effects on the amount dead chlorotic tissue in
shoots and crust regeneration, while higher rates of N addition only negatively impacted
apical meristem growth. In Spain, the abundance of most pigments associated with
cyanobacteria and green algae in soil crusts within Mediterranean ecosystems did not
significantly vary with N deposition along a gradient, but the cyanobacterial pigment
echinenone was negatively correlated with N deposition (4.3-7.3 kg N/ha/vr)(Ochoa-
Hueso et al.. 2016). In China, Zhou et al. (2016) observed that high N addition rates
(50 kg N/ha/yr for 4 years) decreased total chlorophyll and chlorophyll b content in soil
crusts dominated by cyanobacteria. Lower N addition rates (3-15 kg N/ha/yr), however,
did not affect cyanobacteria chlorophyll, and lichen chlorophyll was unaffected by the N
addition treatments. Cyanobacteria soluble sugar concentrations, an osmolyte produced
as a response to stress, increased at the second highest N dose, but other osmolytes
(proline, soluble protein) and lichens were unaffected (Zhou et al.. 2016).
There are very few field studies of how N enrichment impacts belowground C cycling in
arid systems (Liu and Greaver. 2010). In a meta-analysis of these studies, Liu and
Greaver (2010) found no consistent effect of N enrichment on soil respiration. Verburg et
al. (2013) quantified the effects of increased summer precipitation and N deposition on
fine root dynamics in a Mojave Desert ecosystem during a 2-year field experiment.
Increased summer precipitation and 40 kg N/ha/yr N additions did not have an overall
significant effect on any of the measured root parameters. Within the Sonoran Desert near
Phoenix, Marusenko etal. (2015) observed that N additions (60 kg N/ha/yr for 8 years)
increased the abundance of the a mo A gene, which is needed for ammonia oxidation, in
both archaea and bacteria. There was no increase in ammonia-oxidizing bacteria or
archaea along a narrow N deposition gradient (4.3-7.3 kg N/ha/yr) in Mediterranean
shrublands and grasslands in Spain, and overall bacterial abundance was negatively
related to N deposition (Ochoa-Hueso et al.. 2016).Sinsabaugh et al. (2015) conducted a
6-91
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
meta-analysis of soil microbial responses to N additions (5 to 560 kg N/ha/yr; 0.3 to
10 years) in arid ecosystems and identified only 8 studies of various microbial biomass
responses and 10 studies of various microbial metabolic responses. Broadly, both metrics
increased with the addition of low amounts of N, but responses became negative once
threshold N addition rates (kg N/ha/yr: 88 for biomass, 70 for metabolism) or cumulative
loads (kg N/ha: 159 for biomass, 114 for metabolism) had been reached.
In a greenhouse study with California sagebrush (Artemisia californica) seedlings,
Valliere and Allen (2016a) observed that the effects of added N on root colonization by
arbuscular mycorrhizal fungi depended both on whether the soil inoculum came from a
high or low N deposition site and whether the seedlings were stressed by drought; root
colonization was generally lower with soil inoculum from the high N deposition site,
except when N was added under well-watered conditions. Notably, the same was also
true for root colonization by nonmycorrhizal fungi (Valliere and Allen. 2016a). A related
greenhouse study grew a non-native grass and two non-native forbs from CSS
ecosystems in two trials using either sterile soils or soil inoculated from plots receiving N
additions or control plots (Valliere and Allen. 2016b). This study found that N additions
(57 kg N/ha/yr) had no effect on arbuscular mycorrhizal root colonization in one forb
(Hirschfeldia incana), increased arbuscular mycorrhizal colonization in only the second
trial for the other forb (Centaurea melitensis), and increased arbuscular mycorrhizal
colonization in the second trial for the grass (Bromus diandrus) when it was inoculated
with soil from N addition plots (Valliere and Allen. 2016b). Although there were
differences in mycorrhizal colonization between the species and soil types, no direct link
was observed between colonization and the plant growth response to added N; while not
always significant, added N increased aboveground biomass in all plants and in all soil
types. The effect of added N was larger in the second trial and in inoculated soils, but
there was little difference in aboveground growth of the non-native plants between the N
addition inoculum and the control soil inoculum (Valliere and Allen. 2016b). However,
Valliere and Allen (2016a) only observed positive sagebrush seedling growth responses
to added N in soils with high N deposition site inoculum, whereas sterile and low N
deposition inoculum were unresponsive to added N. Thus, evidence is still mixed that soil
microbial communities at sites that have experienced N additions can themselves change
the growth response of plants to added N in arid and semiarid ecosystems.
Vourlitis (2012) measured the aboveground biomass and litter production of a post-fire
chaparral and a mature CSS stand in southern California over an 8-year period. Nitrogen
additions decreased NPP in the chaparral during the first 3 years of the study, but these
additions increased NPP during the last 3 years of the experiment. In the CSS, the effect
of added N positively correlated with precipitation and was only significant in the high
rainfall years. The authors suggested N enrichment (50 kg N/ha/yr) may increase the
6-92
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
productivity, but temporal patterns may take years to emerge and be dependent on water
availability. In contrast, shorter experiments conducted by Vourlitis and Fernandez
(2012) and Vourlitis et al. (2009) in chaparral and CSS ecosystems in southern California
found 4 years of dry season N addition at 50 kg N/ha/yr increased litter N and plant
tissue, but had no effect on ecosystem productivity or ecosystem N storage. Another
study using the same N deposition gradient in the CSS as Padgett and Allen (1999) found
lower soil C content at high deposition sites (Liu and Crowley. 2009). Further,
experimental N additions [50 kg N/ha/yr as Ca(NO;,):| decreased soil C content at the low
deposition site. However, there were no clear changes in microbial community
composition along the deposition gradient or as a result of the N additions.
A study by Vourlitis and Pasquini (2008) analyzed pre- and post-fire nutrient and soil
dynamics of three southern California chaparral stands exposed to varying levels ofN
deposition and determined that periodic fire may not reduce N enrichment from decades
of N deposition. In a subsequent study, (Pasquini and Vourlitis. 2010) exposed chaparral
stands to different levels ofN deposition over the first 3 years of post-fire succession.
High N deposition was associated with a lower relative abundance of A. fasciculatum and
a higher relative abundance of other shrub and herbaceous species. However, overall
aboveground net productivity was not related to N deposition.
Numerous studies were also conducted on dryland ecosystems outside of the U.S.,
including in China. Many of these Chinese N addition experiments have been conducted
in semiarid steppe ecosystems, particularly within the Inner Mongolia region. In these
experiments, N additions increased plant productivity (Zeng et al.. 2016; Zhang et al..
2015d; Li et al.. 2014; Sun et al.. 2014). with a larger stimulation of aboveground growth
than belowground growth (Wang et al.. 2015b; Zhang et al.. 2015d; Li et al.. 2014). In
one of these experiments, 4 years ofN additions (50 kg N/ha/yr as urea) increased NEE
(+53.8%), ER (+47.6%), and GEP (+47.9%) in the last 3 years of the experiment (Wang
et al.. 2015b). The N additions increased foliar N concentrations (Zeng et al.. 2016;
Zhang et al.. 2015a). but Zhang et al. (2015a) also observed that the N additions
decreased the amount ofN resorbed during leaf senescence in seven perennial grass,
sedge, and shrub species. Effects on microbial biomass in these Chinese N addition
experiments have been both positive (Shi et al.. 2016b; Sun et al.. 2014) and negative (Li
et al.. 2016a).
Huang et al. (2015) conducted a 3-year N x water experiment focused on soil microbial
community composition and function in a desert steppe ecosystem in northwestern China
that receives 25 kg N/ha/yr of ambient deposition. The N addition treatment
(50 kg N/ha/yr as NH4NO3) did not have consistent effects, increasing microbial
respiration in the spaces between shrubs in the first year, while decreasing microbial
6-93
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
respiration and microbial biomass C beneath the shrubs in the abnormally dry second
year. The N addition effects differed between inter-plant spaces and beneath shrub sites.
In the inter-plant spaces, N additions tended to increase overall microbial abundance,
with significant increases in bacteria and total microbial PLFA biomass in all 3 years and
an increase in actinobacteria in the last 2 years of the experiment. Beneath the shrubs, N
additions decreased bacterial biomass, fungal biomass, and total PLFA in 2 years and
decreased actinobacterial biomass in all 3 years (Table 6-12).
Zhang et al. (2015e) studied how N added (eight levels from 10 to 500 kg N/ha/yr as
NH4NO3) frequently (12 doses/yr) or infrequently (2 doses/yr) over 5 years influenced
plant productivity in a steppe grassland in the Inner Mongolia region of China. The
higher doses of N increased aboveground NPP regardless of how frequently the N was
added. However, the authors pointed to a group of similar studies that had shown small
frequent N doses caused more, less, or similar growth responses relative to large
infrequent N doses. As in other semiarid systems, the authors noted greater productivity
responses to N in years with higher precipitation. Although the N additions increased
ecosystem aboveground NPP, some individual species were less productive with added N
(Zhang et al.. 20156).
6-94
DRAFT: Do Not Cite or Quote

-------
Table 6-12 Arid and semiarid microbial biomass responses to experimental
nitrogen additions.
Reference
Study
Location
Vegetation
Ambient
Deposition
or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Treseder (2008)
Desert
Meta-analysis
Addition
50-500
0.4-2
Microbial
biomass
Increase
Sinsabauqh et al.
(2015)
Global
Arid lands
meta-analysis
Addition
5-560
0.3-10
Microbial
biomass
Low dose:
increase
Hiqh dose:
decrease
Sinsabauqh et al.
(2015)
Nevada
(Mojave)
Creosote and
bursage
shrublands
(Larrea
tridentata,
Ambrosia
dumosa)
Addition
7, 15
1
Microbial
biomass
Not significant
Sun et al. (2014)
China
(north,
Songnen)
Shrubland
(Leymus
chinensis,
Artemisia
scoparia)
Addition
23, 46,
69, 92
3
Microbial
biomass C
Hiqhest dose:
increase
Other doses:
not significant
Huanq et al. (2015)
China
Desert shrubs
(Haloxylon
ammodendron)
Addition
50
3
Microbial
biomass C
Not significant
Sinsabauqh et al.
(2015)
Global
Arid lands
meta-analysis
Addition
5-560
0.3-10
Microbial
metabolism
Low dose:
increase
Hiqh dose:
decrease
Requs et al. (2017)
California
Desert,
shrubland
Addition
1.8-8.7+
8 weeks
Nodule
formation
by rhizobia
on native
legume
Decrease
(.Acmispon
strigosus)
C = carbon; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints or ecosystems were measured. References ordered by
endpoint. Only statistically significant effects are listed as increases or decreases.
June 2018
6-95
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
In a greenhouse study of plants collected from a semiarid ecosystem in Spain, Ochoa-
Hueso and Manrique (2010) found that grass biomass was relatively unresponsive to N
enrichment of up to 50 kg N/ha, even when provided with increased water. However,
forb biomass and community biomass approximately tripled when both water and N were
enriched. In a related field experiment in central Spain, Ochoa-Hueso and Stevens (2015)
added N (10, 20, or 50 kg N/ha/yr as NH4NO3) to a semiarid shrubland ecosystem for
6 years. The effects on plant productivity varied by time. After 2.5 years, N additions
decreased the biomass of the dominant forb species. After 5.5 years, N additions
increased the biomass of plants in the Cruciferae family, but only in areas where
sufficient soil P was available. The authors noted that each of the biomass assessments
occurred in years with above-average precipitation. In Year 4 of the experiment, Ochoa-
Hueso et al. (2014) surveyed soil fauna (largely arthropods) and observed that the
moderate dose (20 kg N/ha/yr) significantly increased the total abundance of soil fauna
(organisms/g of soil), principally by increasing Collembola populations. This increase
was not significant at other doses and was attributed to soil acidification. At the highest
dose, soil Pauropoda increased (Ochoa-Hueso et al.. 2014).
Elsewhere in Spain, Taboada et al. (2016) conducted a series of experiments to
understand how N deposition influenced herbivous heather beetles (Lochmaea suturalis)
and their predators (beetles and Arachnids) in a semiarid Calluna vulgaris heathland. In
food-choice lab experiments, beetle larvae caused significantly greater shoot weight loss
in plants that had experienced high N addition rates (20, 50, or 56 kg N/ha/yr) than the
low N addition rate (10 kg N/ha/yr) or control, but N addition had no influence on shoot
consumption by adult beetles. Taboada etal. (2016) suggested that this was evidence that
the larvae could detect differences in tissue N and preferred to feed on high-N shoots.
The number of herbaceous heather beetle larvae was greater in plots that had received
long-term N additions (56 kg N/ha/yr for 10 years) than in the control or short-term
(2 year) N addition plots. In contrast, the number of predaceous Arachnids was
unaffected by N additions and the number of predatory beetles was greater in the control
plots, particularly in recently burned plots (Taboada et al.. 2016).
6-96
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
6.3. Relationships between Nitrogen Deposition and Terrestrial
Species Composition, Species Richness, and Biodiversity
6.3.1. Introduction
Given the role of N in biogeochemical cycling (Appendix 4). soil acidification
(Appendix 5). and the growth and physiology of terrestrial organisms (Appendix 6.2). it
is unsurprising that N availability can influence community composition and biodiversity.
In the 2008 ISA, the evidence was sufficient to infer a causal relationship between N
deposition on the alteration of terrestrial species composition, richness, and biodiversity.
In the 2008 ISA, the most sensitive terrestrial taxa were lichens. Empirical evidence
indicated that lichen composition and richness in the U.S. are adversely affected by
deposition levels as low as 3 kg N/ha/yr. Alpine ecosystems were also found to be
sensitive to N deposition. Changes in individual species were estimated to occur at
deposition levels near 4 kg N/ha/yr, and modeling indicated that deposition levels near
10 kg N/ha/yr alter plant community assemblages.
Since 2008, a large number of studies have examined the relationship between N and
terrestrial community composition and species richness. The effects of N deposition on
diversity have been observed in all major biomes (Bobbink et al.. 2010). and these effects
have been particularly well studied in the temperate ecosystems prevalent across much of
the U.S. (Murphy and Romanuk. 2016). Changes in biological communities have been
found across a wide range of taxonomic and functional classifications, including trees,
shrubs, grasses, forbs, bryophytes, fungi (including mycorrhizae), bacteria, and
arthropods. Among these effects, N deposition has been identified as a threat to
12 terrestrial plant or animal species that are listed or candidates for protection under the
federal Endangered Species Act (Hernandez et al.. 2016).
A number of global, continental, and regional syntheses have compiled information about
changes in biodiversity as a result of N deposition. Some of these studies predate the
2008 ISA and served as part of the basis for the conclusions in that assessment
[e.g., (Phoenix et al.. 2006; Pennings et al.. 2005; Suding et al.. 2005; Stevens et al..
2004; Bobbink et al.. 2003; Matson et al.. 2002; Gough et al.. 2000; Wallenda and
Kottke. 1998; Vitousek et al.. 1997)1. More recent assessments have reinforced the
conclusion that N deposition is a major cause of species loss in ecosystems around the
world [e.g., (van Den Berg et al.. 2016; Phoenix et al.. 2012; De Schrijver et al.. 2011;
Pardo etal.. 2011a; Bobbink et al.. 2010)1. Further, these new assessments have provided
a more detailed understanding of both sensitive and heavily impacted species,
ecosystems, and biomes and have created a more precise understanding of the
June 2018
6-97
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
mechanisms linking species loss to N deposition. This new research, combined with
information from the 2008 ISA, creates a body of evidence sufficient to infer a causal
relationship between N deposition and the alteration of species richness, community
composition, and biodiversity in terrestrial ecosystems.
6.3.2. Mechanisms Operating across Terrestrial Ecosystems
Given the complexities of N effects, it is difficult to develop mechanistic models that can
accurately predict shifts in the composition of ecological communities as a result of N
deposition. Consequently, assessments of how N deposition alters species diversity,
species richness, and changes in community composition predominately rely on
experimental manipulations and empirical relationships developed across large spatial
and temporal scales [e.g., (Simkin et al.. 2016; Verheven et al.. 2012; Clark et ak. 2007;
Suding et ak. 2005)1.
Much of the ecological theory regarding the loss of species richness with N enrichment
has been developed from observations of changes in plant species abundance in
grasslands and other nonforest ecosystems [e.g., (Farrcr and Suding. 2016; Pennings et
ak. 2005; Suding et ak. 2005; Gough et ak. 2000)1. Many investigators have found a
unimodal relationship between plant species richness and aboveground primary
productivity, such that there is a negative relationship between productivity and
local-scale plant species richness in terrestrial ecosystems that takes effect once an initial
threshold for minimum productivity has been reached (Suding et ak. 2005; Gough et ak.
2000; Gross et ak. 2000; Grime. 1973). In a synthesis of long-term (>4 years) N addition
experiments (90-130 kg N/ha/yr) in open-canopy ecosystems (grasslands, tundra, etc.),
Gough et ak (2000) found that as the relative stimulation of aboveground primary
productivity by N additions increased, there was a greater relative loss of species
richness. Thus, there was a direct linkage between the stimulation of productivity by
added N and the loss of plant species richness.
Variation among plants in the magnitude of growth responses to added N can have
critical and asymmetric effects on community composition because small differences in
plant height can cause substantial differences in the amount of light absorbed by plant
foliage (Farrer and Suding. 2016). There are a number of potential causes for the
variation in plant response to added N within communities. Two meta-analyses published
in 2005 synthesized data from nonforested N addition experiments across North America
to understand the role of plant functional traits versus random effects in determining
changes in community diversity (Pennings et ak. 2005; Suding et ak. 2005). The
random-loss hypothesis of biodiversity decline suggests that rare species are most likely
June 2018
6-98
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
to disappear as increased competition for resources such as light eliminates less
successful individuals, whereas the functional trait hypothesis predicts that organisms
will become more or less successful when N is added depending upon particular traits.
These mechanisms could operate simultaneously (Sliding et al.. 2005). In this synthesis,
N additions increased primary production and decreased species richness in all terrestrial
ecosystems (Suding et al.. 2005). There was also support for random loss of species: the
rarest species had a >60% chance of disappearance, while the chance of disappearance
for the most abundant species was 10%. This suggests ecosystems with many rare species
that experience increases in competition after N additions are especially sensitive to
species loss.
There was also support for the functional trait hypothesis: N fixing forbs, species shorter
in stature, and perennial plants were more likely to be lost, although there was variation
among the sites in which traits made species the most sensitive to N additions. Pennings
et al. (2005) took a somewhat different approach by following the fate of 20 individual
species across experiments. They noted consistent responses in 10 of the 20 species, but
for all but two species, these responses were dependent on the combination of site
productivity, species richness, functional traits, and initial abundance (Pennings et al..
2005). In a related synthesis of these data to understand the environmental and
experimental factors associated with the loss of species richness, Clark et al. (2007)
observed that several individual factors were almost equally correlated to the loss of
species richness, including aboveground plant production (r = -0.50), lower soil cation
exchange capacity (r = 0.48), and length of the experiment (r = -0.57). Overall, a full
multifactor structural equation model indicated that increased risk of species loss among
the study sites was tied to lower average annual minimum temperature, low soil cation
exchange capacity, and larger productivity responses to N additions (Clark et al.. 2007).
This suggests that high latitude or high elevation ecosystems, those with low cation
exchange capacity (often associated with traits such as sandy soils, low pH, low organic
matter content, and highly weathered soils), and productivity that is strongly limited to N,
would be the most sensitive to plant species loss—traits common in alpine and Arctic
tundra, among other ecosystems.
More recently, De Schriiver et al. (2011) conducted a meta-analysis of plant species
richness responses in N addition experiments conducted in forests (understory plants
only), heathlands, grasslands, tundra, and wetlands. Additions of N increased graminoid
biomass, decreased bryophyte biomass, and decreased the biomass of understory plants in
forests. The addition of N decreased species richness when analyzed across experiments;
within individual ecosystems the effect was only significant in grasslands and heathlands,
while other ecosystems had negative effects but small sample sizes. The loss of species
richness increased with higher cumulative N loads.
6-99
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
The alteration of species interaction by added N occurs via several mechanisms. In
addition to altered interactions between species within the same trophic level
(e.g., competition between plant species), an increased supply of N can change
interactions across trophic levels. These trophic shifts can include changes in the
plant-fungal symbioses that occur in mycorrhizae (Appendix 6.2.3.2 and
Appendix 6.3.3.3). the photobiont-mycobiont relationships in lichens (Appendix 6.2.3.3
and Appendix 6.3.7). and more complex changes in foodwebs that can occur as a
consequence of decreases in plant C allocation to belowground processes
(Appendix 6.2.3.2) and chemical changes in the plant tissues and litter that are consumed
by herbivores and detritivores (Appendix 6.2.3.1. Appendix 6.3.3.4V Changes at higher
trophic levels can feed back to affect the productivity and diversity of primary producers.
For instance, in a greenhouse experiment conducted by Farrerand Suding (2016) using
plants and soils from three long-term N addition experiments (prairie, alpine tundra, arid
grassland), mesocosms inoculated with microorganisms collected from soils that had
received the long-term N additions were significantly more productive and less diverse
than mesocosms inoculated by microorganisms from soils that had not received added N;
these effects occurred regardless of whether the mesocosms received N additions.
Valliere and Allen (2016a) conducted a similar experiment with California sagebrush
(Artemisia californica) and observed that plant productivity was greater under low N
deposition conditions when plants were grown with inoculum from low deposition sites
and greater under high N deposition conditions with inoculum from high deposition sites.
These types of interactions complicate predictions of species loss from N deposition.
A new analysis by Simkin et al. (2016) has dramatically increased our understanding of
how N deposition is altering plant biodiversity in the continental U.S. They gathered data
on herbaceous species richness from more than 15,000 study plots in a variety of habitats.
N deposition had a strong effect on species richness, but this effect differed between
closed canopy and open ecosystems (i.e., forests and nonforest). In nonforested
ecosystems (grasslands, deserts, shrublands, subalpine ecosystems), there was a positive
relationship between N deposition and herbaceous species richness at low rates of N
deposition, then a decrease in species richness with higher rates of N deposition over a
threshold of 8.7 kg N/ha/yr. The effect was notably stronger at low pH. In forests, the
relationship between N deposition and herbaceous species richness was similar, but more
complex. In acidic soils, the relationship was similar to open canopy systems. In neutral
or basic soils, the relationship between N deposition and species richness was
consistently positive. Overall, these national-scale results are consistent with the
relationship between productivity and species richness suggested by ecological theory
[e.g., (Clark et al.. 2007; Suding et al.. 2005)1 and provide evidence that vegetation on
acidic soils is more susceptible to species loss under elevated N.
6-100
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
There are few direct analyses comparing the impacts of oxidized and reduced forms of N
deposition on biodiversity. Because NOs tends to be more readily lost to both leaching
and denitrification than NH4 (Brady and Weil. 1999). it is possible that this form of N is
less likely to accumulate in the soil and cause fewer changes in plant and microbial
communities. The meta-analyses literature referenced previously (Table 6-1) tended to
find no difference in the effects of individual forms of N on ecological and
biogeochemical endpoints, such as plant productivity or microbial biomass. This is
suggestive that plant diversity is also not affected, yet a number of individual studies
have observed differential effects on diversity of NH4+ versus NO3 additions [e.g., (Dias
et al.. 2014; Kleiin et al.. 2008)1. For example, in a nutrient-poor, Mediterranean site, an
NH4+ addition (40 kg N/ha/yr) increased plant richness, while the addition of a half NH44"
and half NO3 mixture (for a total of 40 kg N/ha/yr also) did not. In the U.K., van Den
Berg et al. (2016) observed that once an overall negative effect of N deposition on plant
species richness had been accounted for, the NHx:NOy ratio decreased species richness in
grasslands, but increased species richness in woodlands. By contrast.Jovan et al. (2012)
observed that lichen communities in the Los Angeles basin were best predicted by total N
deposition (as throughfall) than the deposition or gaseous concentrations of any particular
form of N. The national-scale analysis of herbaceous species richness completed by
Simkin et al. (2016) only evaluated total N deposition, not individual forms of deposition.
Future effects of N deposition on biodiversity will depend in part on not only N
emissions, but also on the rate at which populations and communities respond to
increases or decreases in emission rates and the pace at which total cumulative N
deposition increases. Compared to the number of N addition studies, there have been
relatively few studies of recovery following the cessation of N additions or decreases in
N deposition, and nearly all of these have been in Europe (Stevens. 2016). In a review of
these recovery studies, Stevens (2016) found that soil nitrate and ammonium
concentrations recovered to levels observed in untreated controls within 1 to 3 years, but
that soil processes such as N mineralization and litter decomposition were slower to
recover. For instance, differences in decomposition and soil N2O emissions persisted 4
and 7 years, respectively, after installation of a roof system to decrease atmospheric
deposition to European forests (Borken and Beese. 2002; Boxman et al.. 1998b).
Although there were observations that plant physiological processes recovered in less
than 2 years, grassland plant communities were slower to recover and still differed from
controls 11 to 20 years after the cessation of N additions (Isbell et al.. 2013b; Stevens et
al.. 2012). Stevens (2016) observed a wide range of recovery times for mycorrhizal
community composition and abundance, with some recovery observed in as few as
4 years (Emmett et al.. 1998) and differences persisting for as long as 28 or 48 years
(Strengbom et al.. 2001). Based on the slow pace at which plant community composition
recovered following observed or experimental decreases in deposition elsewhere in
6-101
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
Europe, Stevens et al. (2016) predicted that despite future decreases in reactive N
emissions, habitat suitability would decline in 2020 and 2030 for a majority of plant
species within grasslands, heathlands and bogs, and deciduous forests in Great Britain
due to the cumulative effects of N deposition. Plant communities in these ecosystems
were predicted to become more eutrophic: favoring grasses, decreasing the abundance of
forbs and lichens, and having mixed species-specific effects on bryophytes and other
plant groups (Stevens et al.. 2016).
6.3.3. Forests
6.3.3.1.	Overstory Trees
The 2008 ISA reported that there was very little information on the effect of N deposition
on the biodiversity of overstory trees in the U.S. However, the altered growth rates
caused by N enrichment have the potential to affect forest structure and biodiversity. The
life span of many trees is 100 years or more; therefore, it is difficult to observe how
changes in growth or mortality affect biodiversity within established forests. The 2008
ISA cited evidence for boreal forest encroachment into grasslands under increased N
deposition (Kochv and Wilson. 2001). Since the 2008 ISA, little new direct information
is available on the effect of N deposition on the biodiversity of overstory trees in the U.S.
or Europe. In subtropical forests in China along a rural to urban N deposition gradient of
30 to 43 kg N/ha/yr across four sites, Huang et al. (2012) found a decrease in tree
diversity with increasing N deposition, but there were also differences in management
regimes, elevation, and dominant vegetation among the study sites. There is widespread
evidence of species-specific effects of N deposition on tree growth and mortality in the
U.S. [e.g., (Dictzc and Moorcroft. 2011; Thomas et al.. 2010)1. but this information has
not yet been transformed into a quantitative assessment of changes in tree community
composition. A broad-scale analysis of changes in overstory community composition as a
result of N deposition remains a research need.
6.3.3.2.	Understory Plants
Compared to overstory trees, there was more information available in the 2008 ISA
regarding the effects of N deposition on understory plants. A major influence on the 2008
ISA was a review by Gilliam (2006). who identified nine studies on N deposition effects
on forest understory plant communities, all within the U.S. and Europe. All of the North
American studies cited by Gilliam (2006) were N addition experiments within temperate
June 2018
6-102
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
forests in the northeastern U.S.; two of these studies found no significant changes in
community composition (Gilliam et al.. 2006; Gilliam et al.. 1994). Notably, research by
Gilliam et al. (2016b) has subsequently reported decreases in understory species richness,
community evenness, and diversity, and changes in community composition at one of the
northeastern temperate forest sites where understory community composition had earlier
been unaffected by chronic N additions (Gilliam et al.. 2006; Gilliam et al.. 1994). In
particular, the change in understory community composition caused by N additions
included a large increase in the cover of blackberry (Rubus spp.) shrubs (Gilliam et al..
2016b; Walter et al.. 2016). but more broadly included an increase in nitrophilous species
and a loss of more N efficient (oligotrophic) species (Gilliam et al.. 2016b). In Sweden,
shifts in understory plant community composition were documented along two forest N
deposition gradients that ranged from 6 to 20 kg N/ha/yr (Brunet et al.. 1998) and from 3
to 12 kg N/ha/yr (Strengbom et al.. 2001).
Many of the mechanisms through which N additions alter plant community composition
and species richness in other terrestrial ecosystems also function in forest understory
environments, including shifts in mycorrhizal communities, competitive exclusion,
increases in herbivory, and species-specific growth responses to increased N availability
(Gilliam. 2006). However, resource heterogeneity in forests can be especially high, and
limited light availability can constrain the effects of competitive exclusion (Simkin et al..
2016). Since the 2008 ISA, a number of new studies of N addition and ambient
deposition effects on understory plant diversity have been conducted in temperate and
boreal forests in the U.S. and Europe, as well as in subtropical forests in China
(Table 6-13). In a national-scale analysis of herbaceous species richness, Simkin et al.
(2016) found that the effect ofN deposition on forest understory species richness was
highly dependent on soil pH. At all soil pH values, low rates of N deposition
(~5 kg N/ha/yr) had neutral or positive effects on species richness. At a low pH (4.5),
species richness declined at N deposition rates >11.6 kg N/ha/yr, but among neutral and
basic soils there was no point in the data set at which N deposition had a negative effect
on species richness (the analysis included deposition values up to -20 kg N/ha/yr). In a
study that synthesized data from >1,200 forest understory vegetation plots in northern
and central Europe, Verheven et al. (2012) did not find a significant effect of increasing
N deposition (8.3 to 35.7 kg N/ha/yr) on species richness. Although understory plant
species richness did not change, community composition shifted significantly away from
light-demanding species and toward nutrient-demanding species, van Dobben and de
Vries (2010) found a similar result in 366 plots in managed forests throughout western
and northern Europe and in a smaller network of 197 plots within the Netherlands: N
deposition was not a significant influence on plant species richness or diversity and only
a minor influence on community composition via a shift toward more nitrophyllic
species. Soil, elevation, climate, and overstory tree composition were much more
6-103
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
important determinants of community composition at this scale (van Dobben and de
Vries. 2010). In a study of forest understory vegetation plots at 28 sites across Europe,
Dirnbock et al. (2014) also found that although understory plant species richness was not
affected by N deposition, plant communities became more strongly dominated by both
more nutrient-demanding species and more shade-tolerant species as N deposition rates
increased beyond identified critical loads.
There is also evidence for shifts in plant communities across local and regional N
deposition gradients. In a survey of 260 sessile oak (Quercuspetraea) forests in Ireland,
Wilkins and Aherne (2016) observed that N deposition was negatively correlated with
plant species richness and had a significant influence on community composition.
Community composition and species richness were also significantly influenced by soil
pH (positive effect on species richness) and atmospheric NH3 concentrations (negative
effect). Ten species were positively associated with total N deposition and atmospheric
NH3, while 10 other species had negative associations with these N variables; each group
included shrubs, trees, ferns, and bryophytes (Wilkins and Aherne. 2016). The authors
suggested that focusing on a single forest type allowed them to better isolate the effect of
N deposition relative to studies of broader forest communities in Europe (Wilkins and
Aherne. 2016). In a survey of woodlands across the U.K., van Den Berg et al. (2016)
observed a negative effect ofN deposition on plant species richness. Huang etal. (2012)
observed that understory plant species richness was negatively correlated with N
deposition along a rural to urban gradient (30 to 40 kg N/ha/yr) of four subtropical forests
in China. At a network of 40 sugar maple forest monitoring sites in Ontario, McDonough
and Watmough (2015) observed that most of the variability in understory plant
community composition could be explained by climate and soil factors, but that N
deposition (8 to 12.9 kg N/ha/yr) explained a statistically significant portion of the
variability. However, there was no effect on understory species richness over this
relatively narrow range of N deposition.
6-104
DRAFT: Do Not Cite or Quote

-------
Table 6-13 Forest plant diversity responses to nitrogen added via atmospheric deposition or experimental
treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Huanq et al. (2012)
China
(southern
tropical)
Moist subtropical
evergreen broadleaf
forests
Ambient
30.1-43.1
n/a
Overstory
species richness
Decrease
(Walter et al.. 2016)
West Virginia
Blackberry and berries in
the Rubus genus
Addition
35
18
Relative cover
of Rubus spp.
Increase, when N
addition
accompanied by
greater forest
canopy openness
Talhelm et al. (2013)
Michigan
(four sites)
Northern hardwood (Acer
saccharum)
Addition
30
+ 10
Tree sapling
community
composition
Not significant
Talhelm et al. (2013)
Michigan
(four sites)
Northern hardwood (Acer
saccharum)
Addition
30
+ 10
Tree sapling
species richness
Not significant
Strenqbom and Nordin
(2008)
Sweden
Scots pine, Norway
spruce, birch (Pinus
sylvestris, Picea abies,
Betula)
Addition
150 (twice)
Additions 22
and 30 yr
prior to
surveys
Understory
community
composition
Change
Verheven et al. (2012)
Central and
northern
Europe
Temperate deciduous
forests
Ambient
8.3-35.7
n/a
Understory
community
composition
Change
Hedwall et al. (2013)
Sweden
Norway spruce (Picea
abies)
Ambient
4.4-16.1
n/a
Understory
community
composition
Change
June 2018
6-105
DRAFT: Do Not Cite or Quote

-------
Table 6 13 (Continued): Forest plant diversity responses to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Dirnbock et al. (2014)
Europe
Boreal, temperate, and
Mediterranean forests
Ambient
0.6-20.2
n/a
Understory
community
composition
Change
McDonouqh and
Watmouqh (2015)
Ontario,
Canada
Northern hardwood (Acer
saccharum)
Ambient
8.3-12.9
n/a
Understory
community
composition
Change
Chapman et al. (2016)
Pennsylvania
Mixed oak (Quercus)
Addition
100, 200
4
Understory
community
composition
Not significant
Du (2017)
China
(northeastern)
Boreal forest (Larix
gmelinii)
Addition
20, 50, 100
3
Understory
community
composition
Change
Gilliam et al. (2016b) West Virginia Mixed temperate	Addition	35	25 Understory Increase in
hardwood forest	herbaceous understory
cover and	herbaceous
species diversity cover; decrease
in species
diversity
Talhelm et al. (2013)
Michigan Northern hardwood (Acer Addition
(four sites) saccharum)
30
+ 10 Understory plant Change
community
composition
Strenabom and Nordin
(2008)
Sweden	Scots pine, Norway
spruce, birch (Pinus
sylvestris, Picea abies,
Betula)
Addition
150 (twice)
Additions 22
and 30 yr
prior to
surveys
Understory
species diversity
Decrease
Jones and Chapman Pennsylvania Mixed oak (Quercus) Addition	13	1	Understory Not significant
(2011)	species diversity
Talhelm et al. (2013) Michigan Northern hardwood (Acer Addition	30	+10 Understory Not significant
(four sites) saccharum)	species diversity
June 2018	6-106	DRAFT: Do Not Cite or Quote

-------
Table 6 13 (Continued): Forest plant diversity responses to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Lu etal. (2010)
China
(southern
tropical)
Evergreen tropical moist
forest
Addition
50, 100, 150 (35
ambient)
6
Understory
species richness
Decrease
Jones and Chapman
(2011)
Pennsylvania
Mixed oak (Quercus)
Addition
13
1
Understory
species richness
Not significant
Lu etal. (2011c)
China
(southern
tropical)
Pine and broadleaf (Pinus
massoniana, Schima
superba)
Addition
50, 100
(40 ambient)
6
Understory
species richness
Not significant
Huanq etal. (2012)
China
(southern
tropical)
Moist subtropical
evergreen broadleaf
forests
Ambient
30.1-43.1
n/a
Understory
species richness
Decrease
Verheven et al. (2012)
Central and
northern
Europe
Temperate deciduous
forests
Ambient
8.3-35.7
n/a
Understory
species richness
Not significant
Talhelmetal. (2013)
Michigan
(four sites)
Northern hardwood (Acer
saccharum)
Addition
30
+ 10
Understory
species richness
Not significant
Dirnbock et al. (2014)
Europe
Boreal, temperate, and
Mediterranean forests
Ambient
0.6-20.2
n/a
Understory
species richness
Not significant
McDonouah and
Watmouqh (2015)
Ontario,
Canada
Northern hardwood (Acer
saccharum)
Ambient
8.3-12.9
n/a
Understory
species richness
Not significant
Chapman et al. (2016)
Pennsylvania
Mixed oak (Quercus)
Addition
100, 200
4
Understory
species richness
Not significant
Simkin et al. (2016)
Contiguous
U.S.
Forests
Ambient
1.3-17.9
n/a
Understory
species richness
Low soil pH:
decrease
Hiah soil pH:
increase
June 2018
6-107
DRAFT: Do Not Cite or Quote

-------
Table 6 13 (Continued): Forest plant diversity responses to nitrogen added via atmospheric deposition or
experimental treatments.



Ambient
Nitrogen

Effect of

Study

Deposition or
Addition Rate

Additional
Reference
Location
Vegetation
Addition
(kg N/ha/yr)
Duration (yr) Endpoint
Nitrogen
Du (2017)
China
Boreal forest (Larix
Addition
20, 50, 100
3 Understory
Not significant

(northeastern)
gmelinii)


species richness

ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-108
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Understory plant species richness was also not affected in several N addition studies in
temperate forests in the U.S. (Chapman et al.. 2016; Talhelm et al.. 2013; Jones and
Chapman. 2011). In an oak forest in the Philadelphia, PA area, Jones and Chapman
(2011) found that native understory plant diversity and richness were negatively
correlated with soil inorganic N, but not significantly impacted by N additions
(13 kg N/ha/yr). In a follow up at the same site, Chapman et al. (2016) used higher rates
of N additions (100 or 200 kg N/ha/yr as NH4NO3 for 4 years). Changes in total, native,
or invasive species cover were not detected, but N additions increased invasive species
richness in some years. In Michigan, Talhelm et al. (2013) found that although N
additions did not alter understory plant species richness, community composition was
altered. Tree sapling communities were apparently more resistant, as neither species
richness nor community composition changed in response to N additions (Talhelm et al..
2013).
The effects of N additions on forest understory plant communities can be long lasting.
Strengbom and Nordin (2008) found significant differences in understory plant
communities in recently harvested regenerating Swedish forests as a result of two single
additions of 150 kg N/ha 22 and 30 years earlier. The forests that received N decades
earlier had denser understory vegetation, much less shrub cover, and much more herb and
graminoid cover. The abundance of individual species, including ground-dwelling
bryophytes and lichens, differed by as much as 20 times due to the N additions
(Strengbom and Nordin. 2008). In a related study, Hedwall et al. (2013) found that the
amount of change in understory plant community composition in regenerating forests
caused by N additions was negatively correlated with the rate of N deposition along a
Swedish deposition gradient of 4.4 to 16.1 kg N/ha/yr. Wardle et al. (2016) observed that
6 years of N additions (100 kg N/ha/yr) decreased species richness and altered
community composition in boreal forest understory communities dominated by Ericaeous
shrubs and bryophytes.
6.3.3.3. Microbial Diversity
There is an ongoing debate about how to define a microbial species, so taxonomists often
use measures of similarity in genetic material to classify microbial diversity (Fraser et al..
2009). This can lead to discussion of biodiversity changes in terms of taxonomic
community shifts or taxonomic richness. For instance, Turlapati et al. (2013) measured
microbial diversity in surface soils at Harvard Forest by clustering operational taxonomic
units (OTUs) based on 97% sequence similarity from 16S rRNA pyrosequencing.
June 2018
6-109
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Soils contain a high amount of microbial diversity (Lynch et al.. 2012). but this diversity
is not evenly distributed among phylogenetic groups. In the Harvard Forest study,
Turlapati et al. (2013) observed that 2% of the OTUs contained >50% of the total
sequences, while 80% of the OTUs were infrequently observed and contained only about
10% of total sequences. This suggests that many individual soil microbes are either from
the same or closely related taxonomic groups, while concomitantly there are a large
number of relatively rare taxonomic groups. Reanalyzing these data with a more detailed
oligotype sequence clustering approach, Turlapati et al. (2015) found a more even
distribution of taxonomic groups, but 2% of oligotypes still represented -38% of the total
number of sequences. In an oligotype analysis of soil fungal communities at the same
site, Morrison et al. (2016) observed a similar distribution of fungal populations: the
Basidiomycota phylum comprised 63-71% of all sequences, while the Ascomycota
phylum averaged 26%, and the ectomycorrhizal genus Russula made up 40-50% of all
sequences across plots. A shotgun metagenomics assessment of microbial communities in
four northern hardwood forests in Michigan documented that taxonomic classification of
the metagenomics hits via the RDP databased within MG-RAST were heavily dominated
by bacteria (98% of reads), with much smaller contributions of fungi (1%) and archaea
(0.03%) (Freedman et al.. 2016). However, annotation databases are biased toward
culturable bacteria and there is transcriptomic evidence that metagenomic approaches
greatly underrepresent the fungal contribution to soil metabolic function (Freedman et al..
2016).
Much of the available information about changes in microbial community composition
published since 2008 has been about changes in fungal communities, including
mycorrhizal species. Of the studies identified for this assessment that quantified the
compositional response of forest fungal communities to N additions, one study had
site-specific results, but N additions changed community composition in five of the
remaining seven studies (Table 6-14). Among studies of ectomycorrhizal fungi
community composition, N caused changes in six out of seven studies, including in four
studies along ambient N deposition gradients (Table 6-15). As with lichens and plants,
the sensitivity of ectomycorrhizal fungi appears to vary taxonomically, with these
taxonomic differences apparently related to differences in functional traits such as
organic N acquisition (Lilleskov et al.. 2011). At Harvard Forest, Morrison et al. (2016)
observed that N additions (50 or 150 kg N/ha/yr) decreased the relative abundance of the
ectomycorrhizal genus Cenococcum and increased the ectomycorrhizal genera
Scleroderma and Rhizoscyphus. Scleroderma is thought to be important in organic P
acquisition, and Rhizoscyphus has proteolytic, cellulytic, and partial lignin degradation
capabilities, but less is known about Cenococcum. The overall abundance of the genus
Russula did not change as a result of N additions, but the Russula species (R. vinacea)
6-110
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
that was the single most abundant OTU in the data set increased from 10% relative
abundance to 28-37% in the N addition plots, while the other Russula decreased.
Shifts in ectomycorrhizal composition have also been observed along N depositional
gradients. Kjolleretal. (2012) observed changes in the ectomycorrhizal community
composition and a loss of ectomycorrhizal species richness across an N deposition
gradient in Denmark from 27 to 43 kg N/ha/yr at the edge of a spruce forest. In Scotland,
Jarvis et al. (2013) analyzed ectomycorrhizal fungal communities from 15 seminatural
Scots pine forests and found changes in abundance of Cortinarius spp. at higher N
deposition (9.8 kg N/ha/yr). In red spruce-dominated forests in the northeastern U.S.,
Lilleskov et al. (2008) observed changes in ectomycorrhizal communities over a much
lower N deposition gradient, ranging from 2.8 to 7.9 kg N/ha/yr of wet deposition.
Nitrogen deposition was positively related to fine root N concentrations, and those root N
concentrations had positive relationships with the abundance of three of the
ectomycorrhizal fungal morphotypes, negative relationships with three morphotypes, and
ambiguous relationships with three other morphotypes. In Scots pine (Pinus sylvestris)
forests in Germany and the U.K. arrayed along an N deposition gradient ranging from 4.6
to 28.6 kg N/ha/yr, Cox et al. (2010) used DNA extracted from fine roots and group
sequences at 97% similarity to assess changes in ectomycorrhizal community
composition. Similar to the results in the Lilleskov et al. (2008) study, N deposition was
positively correlated to foliar N concentrations and foliar N was linked to shifts in
mycorrhizal community structure. In particular, foliar N concentrations were significantly
negatively correlated with ectomycorrhizal richness. Of the 35 taxathat occurred widely
across the gradient, 11 showed significant responses to increased N availability: 6 taxa
increased and 5 taxa decreased with greater N availability (Cox et al.. 2010). Thus, the
effects of N deposition on ectomycorrhizal community composition observed by
Lilleskov et al. (2008) and Cox et al. (2010) were indirect.
6-111
DRAFT: Do Not Cite or Quote

-------
Table 6-14 Forest microbial biodiversity responses to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study Location
Ambient
Deposition or
Vegetation	Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Eisenlord et al. (2013)
Michigan (Ml
gradient)
Northern	Addition
hardwood forests
(Acer
saccharum)
30
16 Actinobacterial
community
composition
Three sites: change
One site: not
significant
Eisenlord et al. (2013)
Michigan (Ml
gradient)
Northern	Addition
hardwood forests
(Acer
saccharum)
30
16 Actinobacterial gene Two sites: decrease
diversity	Two sites: not signific;
Eisenlord et al. (2013)
Michigan (Ml
gradient)
Northern	Addition
hardwood forests
(Acer
saccharum)
30
16 Actinobacterial gene Two sites: decrease
functional richness Two sites: not signific:
Krumins et al. (2009)
Florida and New
Jersey
Scrub oak	Addition
forests (Quercus
myrtifolia,
Q. ilicifolia)
35, 70
Bacterial community Not significant
composition
Turlapati et al. (2013)
Massachusetts
(Harvard Forest)
Temperate oak Addition
forest (Quercus
rubra,
Q. velutina)
50,150	22 Bacterial community Change
composition
Hesse et al. (2015)
Michigan (Ml
gradient)
Northern	Addition
hardwood forests
(Acer
saccharum)
30
16 Bacterial community Not significant
composition
June 2018
6-112
DRAFT: Do Not Cite or Quote

-------
Table 6-14 (Continued): Forest microbial biodiversity responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Turlaoati et al. (2015)
Massachusetts
(Harvard Forest)
Temperate oak
forest (Quercus
rubra,
Q. velutina)
Addition
50, 150
22
Bacterial community
composition
Change
Turlapati et al. (2013)
Massachusetts
(Harvard Forest)
Temperate oak
forest (Quercus
rubra,
Q. velutina)
Addition
50, 150
22
Bacterial richness
Increase
Hesse et al. (2015)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
16
Bacterial richness
Not significant
Zechmeister-Boltenstern et
al. (2011)
Europe
Conifer and
broadleaf forests
Ambient
2-40
n/a
Bacterial/fungi ratio
Increase
Allison et al. (2008)
Alaska
Boreal forest
(Picea mariana)
Addition
140
5
Fungal community
composition
Change
Krumins et al. (2009)
Florida and New
Jersey
Scrub oak
forests (Quercus
myrtifolia,
Q. ilicifolia)
Addition
35, 70
1
Fungal community
composition
Not significant
Allison et al. (2010)
Alaska
Boreal forest
(Picea mariana,
Festuca altaica)
Addition
114
7
Fungal community
composition
Change
Edwards et al. (2011)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
14
Fungal community
composition
Not significant
June 2018
6-113
DRAFT: Do Not Cite or Quote

-------
Table 6-14 (Continued): Forest microbial biodiversity responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Eisenlord et al. (2013)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
16
Fungal community
composition
Two sites: chanae
Two sites: not
significant
Hesse et al. (2015)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
16
Fungal community
composition
Change
Gillet et al. (2010)
Switzerland
Norway spruce
(Picea abies)
Addition
150
12
Fungal community
composition
(saprobic fungi)
Change
Allison et al. (2008)
Alaska
Boreal forest
(Picea mariana)
Addition
140
5
Fungal diversity
Not significant
Eisenlord et al. (2013)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
16
Fungal gene
diversity
Two sites: decrease
Two sites: not siqnific;
Eisenlord et al. (2013)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
16
Fungal gene
functional richness
Two sites: decrease
Two sites: not siqnific;
Krumins et al. (2009)
Florida and New
Jersey
Scrub oak
forests (Quercus
myrtifolia,
Q. ilicifolia)
Addition
35, 70
1
Fungal morphotype
richness
Not significant
Hesse et al. (2015)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
16
Fungal richness
Not significant
June 2018
6-114
DRAFT: Do Not Cite or Quote

-------
Table 6-14 (Continued): Forest microbial biodiversity responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Gillet etal. (2010)
Switzerland
Norway spruce
(Picea abies)
Addition
150
12
Fungal richness
(saprobic fungi)
Decrease
van Diepen etal. (2010)
Michigan (Ml
gradient)
Northern
hardwood forests
(Acer
saccharum)
Addition
30
12
Microbial community
composition
Change
Zechmeister-Boltenstern et
al. (2011)
Europe
Conifer and
broadleaf forests
Ambient
2-40
n/a
Microbial community
composition
Change
Hobbieet al. (2012)
Minnesota (Cedar
Creek)
Oak and pine
forests (Quercus
ellipsoidalis,
Pinus strobus)
Addition
100
5
Microbial community
composition
Change
Zhao etal. (2014a)
China (Tibetan
Plateau)
Spruce-fir (Picea
asperata, Abies
faxoniana)
Addition
250
4
Microbial community
composition
Change
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases or
decreases.
June 2018
6-115
DRAFT: Do Not Cite or Quote

-------
Table 6-15 Ectomycorrhizal biodiversity responses to nitrogen added via atmospheric deposition or
experimental N additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen Addition
Rate (kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Avis et al.
(2008)
Illinois
Oak (Quercus alba,
Q. rubra)
Addition
21
4
Community
composition
Change
Lilleskov et
al. (2008)
Northeastern U.S.
Red spruce (Picea
rubens)
Ambient
2.8-7.9 (wet only)
n/a
Community
composition
Change (indirect)
Wriqht et
al. (2009)
British Columbia,
Canada
Western hemlock
(Tsuga
heterophylla)
Addition
300 (once)
7 yr recovery
Community
composition
Not significant
Cox et al.
(2010)
Germany, U.K.
Scots pine (Pinus
sylvestris)
Ambient
4.6-28.6
n/a
Community
composition
Change (indirect)
Ki0ller et
al. (2012)
Denmark
Norway spruce
(Picea abies)
Ambient
27-43
n/a
Community
composition
Change
Jarvis et al.
(2013)
Scotland, U.K.
Scots pine (Pinus
sylvestris)
Ambient
3.1-9.9
n/a
Community
composition
Change
Gillet et al.
(2010)
Switzerland
Norway spruce
(Picea abies)
Addition
150
12
Community
composition
(sporocarps)
Change
Suz et al.
(2014)
Europe (nine
countries)
Oak (Quercus
robur, Q. petraea)
Ambient
5.1-35.5
n/a
Community
evenness
Decrease
Krumins et
al. (2009)
Florida and New
Jersey
Scrub oak forest
(Quercus myrtifolia,
Q. ilicifolia)
Addition
35, 70
1
Morphotype
richness
Not significant
June 2018
6-116
DRAFT: Do Not Cite or Quote

-------
Table 6-15 (Continued): Ectomycorrhizal biodiversity responses to nitrogen added via atmospheric deposition or
experimental N additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen Addition
Rate (kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Avis et al.
(2008)
Illinois
Oak (Quercus alba,
Q. rubra)
Addition
21
4
Species richness
Decrease
Wriaht et
al. (2009)
British Columbia,
Canada
Western hemlock
(Tsuga
heterophylla)
Addition
300 (once)
7 yr recovery
Species richness
Not significant
Ki0ller et
al. (2012)
Denmark
Norway spruce
(Picea abies)
Ambient
27-43
n/a
Species richness
Decrease
Jarvis et al.
(2013)
Scotland, U.K.
Scots pine (Pinus
sylvestris)
Ambient
3.1-9.9
n/a
Species richness
Not significant
Suz et al.
(2014)
Europe (nine
countries)
Oak (Quercus
robur, Q. petraea)
Ambient
5.1-35.5
n/a
Species richness
Decrease
Gillet et al.
(2010)
Switzerland
Norway spruce
(Picea abies)
Addition
150
12
Sporocarp richness
Decrease
Hasselauist Sweden
and
Hoqberq
(2014)
Scots pine (Pinus
sylvestris)
Addition
35, 70
40, 2 yr recovery
for 70 kg treatment
Sporocarp richness
Decrease
Hasselauist Sweden
and
Hoqberq
(2014)
Scots pine (Pinus
sylvestris)
Addition
110
20; 15 yr recovery
Sporocarp richness
Not significant
Hasselquist Sweden
and
Hoqberq
(2014)
Scots pine (Pinus
sylvestris)
Addition
20, 100
6
Sporocarp richness
Not significant at
20; decrease at
100
June 2018
6-117
DRAFT: Do Not Cite or Quote

-------
Table 6-15 (Continued): Ectomycorrhizal biodiversity responses to nitrogen added via atmospheric deposition or
experimental N additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen Addition
Rate (kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Cox et al.
(2010)
Germany, U.K.
Scots pine (Pinus
sylvestris)
Ambient
4.6-28.6
n/a
Taxonomic
richness
Decrease (indirect)
Ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-118
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Although shifts in fungal community composition are widely reported, these shifts are
not necessarily consistent, even in similar ecosystems. Allison et al. (2010) conducted an
N addition experiment in a recently burned boreal forest in Alaska. In that site, the forest
microbial community was dominated by fungi in the Ascomycota, and N additions
increased their abundance. In contrast, Allison et al. (2008) conducted a similar
experiment in a mature boreal forest in Alaska and observed that the microbial
community was dominated by Basidiomycota, and N additions did not change the overall
abundance of fungi in this phylum. However, in both ecosystems, N additions greatly
changed the relative abundance of individual taxa (orders and families) within these
phyla (Allison et al.. 2010; Allison et al.. 2008). Individual components of microbial
communities also appear to react to N additions at different speeds. In a long-term forest
N addition in Switzerland, Gillet et al. (2010) observed that both ectomycorrhizal and
saprotrophic fungal communities responded to an increase in soil N input, but the
ectomycorrhizal community rapidly decreased in species richness, whereas the
saprotrophic community was less affected. The response was highly species specific,
especially for the saprotrophic community. At Harvard Forest, Frev et al. (2004) found
fungal biomass was reduced in fertilized versus control plots, and this was accompanied
by a decrease in phenol oxidase activity, an enzyme produced by saprophytic, white-rot
fungi to decompose lignin. Among fungal saprotrophs at Harvard Forest, Morrison et al.
(2016) found that high rates of N addition (150 kg N/ha/yr for 25 years) decreased the
abundance of the saprotrophic basidiomycete genus Agaricus, which has lignolytic
abilities, and increased the abundance of two saprophytic ascomycete genera (Hypocrea,
Phialophora) that are cellulolytic. These changes in the fungal community at Harvard
Forest are consistent with observations elsewhere (Freedman et al.. 2016; Edwards et al..
2011) that N additions can stimulate cellulose decomposition and inhibit fungal
decomposition of lignin (see Appendix 4).
Decreases in taxonomic richness among fungi appear to be less common than changes in
community composition. Taxonomic richness is often used as a measure of biodiversity
for microbial communities since it can be difficult to define species as noted previously.
Only three studies of taxonomic richness in overall fungal communities were identified
for this assessment (Table 6-14). One study saw changes at two of four study sites
(Eisenlord et al.. 2013); there was no effect in another study (Krumins et al.. 2009); and
Morrison et al. (2016) observed a shift in composition under a very high dose
(150 kg N/ha/yr), but not a more moderate dose (50 kg N/ha/yr). Of the 11 identified
ectomycorrhizal richness studies, decreases were observed in 6 studies (Table 6-15).
Compared to ectomyorrhizal fungi, there is also less information about how N additions
alter arbuscular mycorrhizal and bacterial community composition (Table 6-14.
Table 6-16). Two studies on arbuscular mycorrhizae were identified, one in Ecuador
6-119
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
(Camcnzind et al.. 2014) that observed no effect of added N (50 kg N/ha/yr for 3 years)
on community composition and another in northern hardwood forests in Michigan where
the effect of added N (30 kg N/ha/yr for +12 years) on community composition was
dependent on the measurement technique (van Diepen et al.. 2013; van Diepen et al..
2011). At the Harvard Forest long-term N addition study in Massachusetts (50 and
150 kg N/ha/yr), (Turlapati et al.. 2015; Turlapati et al.. 2013) used molecular techniques
to assess changes in bacterial community composition. Using a 97% sequence identity
approach to clustering OTUs, (Turlapati et al.. 2013) found the chronic N additions
caused major changes in the bacterial phyla of Acidobacteria, Proteobacteria, and
Verrucomicrobia. Using the more detailed oligotype sequence clustering approach,
Turlapati et al. (2015) observed five genera that exclusively appeared in N treated soils
(Aquabacterium, Nitrosospira, Yersinia, Legionella, and Niabella) and eight genera that
were present only in the control plots (Comamonas, Microbacterium, Mycetocola,
Brochothrix, Flavobacterium, Pedobacter, Sphingobacterium, and Terrimonas).
However, the N addition treatments also caused shifts in community structure within
most genera. As with Fierer et al. (2012). microbial communities in the mineral soil were
less affected by the N additions than communities in the organic soil.
6-120
DRAFT: Do Not Cite or Quote

-------
Table 6-16 Arbuscular mycorrhizal responses to nitrogen added via atmospheric deposition or experimental
treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
van Diepen etal. (2011)
Michigan (four
sites)
Sugar maple (Acer
saccharum)
Addition
30
13
Community
composition
Change
van Diepen etal. (2013)
Michigan (four
sites)
Sugar maple (Acer
saccharum)
Addition
30
14
Community
composition
Not significant
Camenzind etal. (2014)
Ecuador
Evergreen tropical
forest
Addition
50
3
Community
composition
Not significant
Chen etal. (2014)
China
Steppe grassland
Addition
100
6
Plant-associated
microbial
phylotype
diversity
Not significant
Chen etal. (2014)
China
Steppe grassland
Addition
100
6
Plant-associated
microbial
phylotype
richness
Not significant
Chen etal. (2014)
China
Steppe grassland
Addition
100
6
Soil microbial
phylotype
diversity
Not significant
Chen etal. (2014)
China
Steppe grassland
Addition
100
6
Soil microbial
phylotype
richness
Decrease
van Diepen etal. (2011)
Michigan (four
sites)
Sugar maple (Acer
saccharum)
Addition
30
13
Taxonomic
diversity
Three sites: not
significant
One site: decrease
June 2018
6-121
DRAFT: Do Not Cite or Quote

-------
Table 6 16 (Continued): Arbuscular mycorrhizal responses to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
van Diepen etal. (2013)
Michigan (four
sites)
Sugar maple (Acer
saccharum)
Addition
30
14
Taxonomic
richness
Decrease
Camenzind etal. (2014)
Ecuador
Evergreen tropical
forest
Addition
50
3
Taxonomic
richness
Decrease
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-122
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Shifts in microbial community composition can alter ecosystem processes and plant
community diversity, in addition to affecting plant productivity as noted previously
(Appendix 6.2.3). Lower microbial diversity has been linked to slower decomposition of
plant litter and can affect, for example, ecosystem C and nutrient cycling (Bardgctt and
van Per Putten. 2014; Handa et al.. 2014). In a global scale studv.Dclgado-Baqucrizo et
al. (2016) found a positive correlation between microbial diversity and a broad suite of
ecosystem services and functions. Mycorrhizal diversity has been shown experimentally
to affect plant diversity. For instance, Van DerHeiiden etal. (1998) concluded that plant
species composition fluctuated greatly under low arbuscular mycorrhizal diversity,
whereas it increased and exhibited greater stability alongside higher mycorrhizal
diversity. Specifically on N effects, Van Per Heiiden et al. (2008) observed smaller
impacts on plant community composition from N additions (100 kg N/ha/yr) when plant
mesocosms were inoculated with arbuscular mycorrhizal fungi than in those without
fungal inoculum. Plant communities in the inoculated mesocosms had greater evenness
among functional groups. Altogether, these studies suggest microbial, including
mycorrhizal, diversity have a positive effect on ecosystem processes and plant
community diversity. This implies that N induced decreases in microbial diversity may in
turn negatively affect these endpoints.
6.3.3.4. Arthropod and Other Invertebrate Diversity
Arthropods can be key components of forest productivity and nutrient cycling because
they can feed on living plant tissues, plant litter, or on litter-degrading fungi, and
arthropod communities can be directly or indirectly altered by changes in plant
productivity and chemistry (Gan et al.. 2014; Throop and Lerdau. 2004). Consequently,
recent research has quantified the response of arthropods to added N (Table 6-17). Both
before and after the 2008 ISA, Jones and colleagues published a series of studies on how
N air pollution altered plant-associated insect communities in mixed conifer forests
outside Los Angeles, CA (Jones etal.. 2011; Jones et al.. 2008; Jones and Paine. 2006;
Jones et al.. 2004). Jones et al. (2004) observed significantly higher tree mortality and
bark beetle activity at a high versus a low N pollution site. Moreover, experimental N
additions at the low pollution site increased mortality and bark beetle activity relative to
unfertilized control plots. In contrast, the opposite occurred at the high pollution site, with
the control plots exhibiting higher tree mortality and beetle activity. The authors
suggested this could be because N deposition already exceeded biological demand at the
high pollution site, with additional N no longer increasing tree mortality. Jones et al.
(2008) found that insect herbivore communities on California black oak (Quercus
kelloggii) trees did not change in response to several years of N additions
June 2018
6-123
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
(150 kg N/ha/yr) at a relatively unpolluted and dry site, but were altered by N additions at
a wetter site that received more ambient N deposition. Jones etal. (2011) found a similar
response for insect herbivore communities associated with bracken fern plants at these
sites, with no change in insect taxonomic richness at the dry and low-deposition site and
increased richness at the wetter, higher deposition site. Notably, as with plant
productivity responses in grassland and arid environments, the effects of added N on
insect abundance and diversity appeared to be strongly dependent on climate (Jones ct al..
2011).
Gan etal. (2013) and Gan et al. (2014) quantified changes in the abundance and
community composition of soil microarthropods and trophic position of soil-dwelling
oribatid mites at the same four northern hardwoods forests in which Talhelm et al. (2013)
quantified changes in understory plant composition. The overall abundance of
microarthropods declined by approximately 45% in response to N additions, a change
that was attributed to changes in soil food webs as a consequence of previously
documented decreases in litter decomposition (Zak et al.. 2008). decreases in mycorrhizal
productivity (van Diepen et al.. 2010). and shifts in the microbial community
composition (van Diepen et al.. 2013; Edwards et al.. 2011; van Diepen et al.. 2011).
More specifically, Gan etal. (2013) observed decreases in two orders of detritivores
(Oribatida, Collembola) and one order of predaceous mites (Mesostigmata), with the
largest decline in the oribatid mites. The decrease in oribatid mites did not affect species
richness or their trophic position, but did cause a shift in the community composition
(Ganetal.. 2014. 2013).
Lastly, in addition to arthropods, other soil invertebrates, such as earthworms or
nematodes, can be affected by N additions (Table 6-17). Romanowicz and Zak (2017).
for example, observed that the abundance of middens created by the non-native
earthworm species Lumbricus terrestris was 363% higher in plots receiving N additions
at one of the four sites in the study. Many northern temperate forests in the U.S. have no
native earthworms and the spread of these non-native worms can cause decreases in soil
C storage, a redistribution of soil C and N as earthworms consume the soil organic
horizon, shifts in soil foodwebs, and changes in understory plant communities (Bohlen et
al.. 2004).
6-124
DRAFT: Do Not Cite or Quote

-------
Table 6-17 Arthropod and other invertebrate responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen Addition
Rate (kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Zehnder and
Hunter (2008)
Georgia
Milkweed
(.Asclepias
tuberosa)
Addition
25, 40
0.1
Aphid carrying
capacity
Increase
Zehnder and
Hunter (2008)
Georgia
Milkweed
(.Asclepias
tuberosa)
Addition
25, 40
0.1
Aphid population
growth rate
Increase
Cha et al. (2010)
Pennsylvania
Northern red oak
(Quercus rubra)
Addition
200
1
Chewing insect
herb ivory
Increase
Romanowicz and
Zak (2017)
Michigan
Northern hardwood
forests (Acer
saccharum)
Addition
30
22
Earthworm
abundance
(Lumbricus
terrestris middens)
Increase
in one of
two sites
where L.
terrestris
was
present
Pavne et al. (2012)
U.K.
Heathland (Calluna
vulgaris)
Addition
10, 20, 40, 80, 120
11, 21
Enchytraeid worm
abundance
Not
significant
Cha et al. (2010)
Pennsylvania
Northern red oak
(Quercus rubra)
Addition
200
1
Galling insect
herbivory
Not
significant
Wissinqer et al.
(2014)
California (Mojave)
Creosote and
bursage shrublands
(Larrea tridentata,
Ambrosia dumosa)
Ambient
2-12
n/a
Harvester ant
(Messor pergandei)
nest density
Increase
June 2018
6-125
DRAFT: Do Not Cite or Quote

-------
Table 6 17 (Continued): Arthropod and other invertebrate responses to experimental nitrogen additions.



Ambient



Effect of



Deposition or
Nitrogen Addition


Additional
Reference
Study Location
Vegetation
Addition
Rate (kg N/ha/yr)
Duration (yr)
Endpoint
Nitrogen
Wissinqer et al.
California (Mojave)
Creosote and
Ambient
2-12
n/a
Harvester ant
Decrease
(2014)

bursage shrublands



(Messor pergandei)



(Larrea tridentata,



nest size



Ambrosia dumosa)





Jones et al. (2011)
California (southern
Bracken fern
Addition
150
3
Insect herbivore
Low dec

CA, two sites)
{Pteridium



abundance
site: not


aquilinum) in mixed




significant


conifer forests




Hiqh dec







site:







decrease
Jones et al. (2008)
California (southern
California black oak
Addition
150
3-4
Insect herbivore
Low deD

CA, two sites)
(Quercus kelloggii)



community
site: not


in mixed conifer



composition
significant


forests




Hiqh dec







site:







change
Jones et al. (2011)
California (southern
Bracken fern
Addition
150
3
Insect herbivore
Low dec

CA, two sites)
{Pteridium



taxonomic richness
site: not


aquilinum) in mixed




significant


conifer forests




Hiqh dec







site:







increase
Eisenhauer et al.
Minnesota (Cedar
Prairie C3 and C4
Addition
40
14
Mesofauna
Not
(2013)
Creek)
grasses, forbs,



abundance
significant


legumes



(arthropods—five







orders)

Eisenhauer et al.
Minnesota (Cedar
Prairie C3 and C4
Addition
40
14
Mesofauna
Not
(2013)
Creek)
grasses, forbs,



diversity
significant


legumes



(arthropods—five







orders)

June 2018
6-126
DRAFT: Do Not Cite or Quote

-------
Table 6 17 (Continued): Arthropod and other invertebrate responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen Addition
Rate (kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Eisenhauer et al.
(2012)
Minnesota (Cedar
Creek)
Prairie C3 and C4
grasses, forbs,
legumes
Addition
40
14
Microarthropod
richness
Decrease
Eisenhauer et al.
(2013)
Minnesota (Cedar
Creek)
Prairie
Addition
40
14
Nematode
abundance
Three
quilds: not
significant
One auild:
increase
One auild:
decrease
Cesarz et al. (2015)
Minnesota (Cedar
Creek)
Prairie C3 and C4
grasses, forbs,
legumes
Addition
40
14
Nematode
community
composition
Change
Eisenhauer et al.
(2013)
Minnesota (Cedar
Creek)
Prairie C3 and C4
grasses, forbs,
legumes
Addition
40
14
Nematode richness
Not
significant
BishoD et al. (2010)
Washington (Mt. St.
Helens)
Primary
successional alpine
meadow
Addition
78
5
Orthoptera
abundance
Increase
Gan et al. (2013)
Michigan (four
sites)
Northern
hardwoods forests
(Acer saccharum)
Addition
30
18
Soil oribatid mite
abundance
Decrease
Gan et al. (2013)
Michigan (four
sites)
Northern
hardwoods forests
(Acer saccharum)
Addition
30
18
Soil oribatid mite
community
composition
Change
Gan et al. (2013)
Michigan (four
sites)
Northern
hardwoods forests
(Acer saccharum)
Addition
30
18
Soil oribatid mite
species richness
Not
significant
June 2018
6-127
DRAFT: Do Not Cite or Quote

-------
Table 6 17 (Continued): Arthropod and other invertebrate responses to experimental nitrogen additions.
Reference
Study Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen Addition
Rate (kg N/ha/yr)
Duration (yr)
Endpoint
Effect of
Additional
Nitrogen
Gan etal. (2014)
Michigan (four
Northern
Addition
30
18
Soil oribatid mite
Not

sites)
hardwoods forests



trophic position
significant


(Acer saccharum)





Bishop et al. (2010)
Washington (Mt. St.
Helens)
Primary
successional alpine
meadow
Addition
78
3
Total arthropod
abundance
Not
significant
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-128
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
6.3.4.
Alpine and Arctic Tundra
There was consistent and coherent evidence in the 2008 ISA indicating that alpine plant
communities are among the terrestrial communities most sensitive to atmospheric N
deposition. The previous assessment identified a number of factors that made these
ecosystems sensitive to N deposition, including low rates of primary production, short
growing seasons, low temperature, and low rates of N mineralization (Bowman and Fisk.
2001; Bowman and Steltzer. 1998; Fisk et al.. 1998; Bowman. 1994; Bowman et al..
1993). Alpine plants are broadly N limited, and increased N inputs have been observed to
cause changes in alpine growth and species composition (Bowman et al.. 2006; Bowman
and Steltzer. 1998; Vitousek et al.. 1997). Alpine plant communities have also developed
under conditions of low nutrient supply, in part because soil-forming processes are poorly
developed, and this was also thought to contribute to their N sensitivity.
Many of the alpine tundra ecosystems in the U.S. are found in the western U.S.,
particularly in the Rocky Mountains, Cascade Mountains, and the Sierra Nevada.
Atmospheric deposition tends to be low in the western U.S., but alpine sites can be
affected by relatively low or moderate levels of N deposition because of their high
sensitivity. In several studies, the capacity of Colorado Rocky Mountain alpine
catchments to sequester N was exceeded at input levels of 10 kg N/ha/yr or less
[e.g., (Baron et al.. 1994)1. Similarly, relatively low N addition rates can cause shifts in
alpine plant communities. Bowman et al. (2006) estimated that changes in the abundance
of Carex rupestris occurred at deposition rates of 4 kg N/ha/yr and changes in alpine
tundra community composition occurred at deposition rates of approximately
10 kg N/ha/yr. In Europe, critical loads for alpine plant communities were estimated to be
between 5 and 15 kg N/ha/vr (Bobbink et al.. 2003).
Since 2008, there was further research in the U.S. on the effects of added N on plant
communities in Colorado at the Niwot Ridge Long-Term Ecological Research Site and in
Rocky Mountain National Park (RMNP), as well as on Mount St. Helens in Washington
(Table 6-18). In RMNP, Bowman et al. (2012) conducted a 4-year study in a dry alpine
meadow ecosystem where N was added at rates of 5, 10, and 30 kg N/ha/yr, while
ambient N deposition was 4 kg N/ha/yr. No shifts in species richness or diversity were
observed in response to the N additions; however, Carex rupestris increased in cover
from 34 to 125% in response to the treatments(Bowman et al.. 2014). More broadly, as
noted in Appendix 5. McDonnell et al. (2014a) used the ForSAFE-VEG model to assess
changes in vegetation cover in subalpine ecosystems in RMNP since 1900 and forecast
the changes through the end of the 21st century as a result of N deposition and climate
change scenarios. Based on the model output, plant community composition changed by
June 2018
6-129
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
10% over the past 100 years as a result of increases in N deposition, with increases in
graminoid species and decreases in forb species. Over the next 100 years, forecasted
changes in N deposition and climate factors are predicted to increase tree cover. At Niwot
Ridge, two decades of N additions (averaging -75 kg N/ha/yr) generally decreased plant
species richness across a wet-to-dry gradient of alpine meadows, with a significant
decrease in the moist meadow type (Yuan et al.. 2016). Plant diversity increased with N
additions (78 kg N/ha/yr for 5 years) on Mount St. Helens, but this plant community was
growing on very N poor soils formed following the volcanic eruption there in 1980
(Bishop et al.. 2010).
Internationally, plant diversity research has been conducted in Arctic ecosystems in
Scandinavia and North Atlantic Europe (Greenland, Iceland, northern Great Britain, etc.)
and in alpine ecosystems in Switzerland and China. In a subalpine grassland in the Swiss
Alps, Bassin et al. (2013) observed that 7 years of N additions of 5 to 50 kg N/ha/yr
decreased plant diversity and changed plant community composition. Arens et al. (2008)
found no effects of N additions (5, 10, or 50 kg N/ha/yr) on plant community
composition over 3 years in dwarf shrub/herb tundra ecosystems in Greenland. In
Sweden, Sundqvist et al. (2014) did not observe any change in plant species diversity or
plant species richness as a result of 3 years of N additions (100 kg N/ha/yr) to a tundra
heath and a tundra meadow. Elsewhere in Sweden, Wardle etal. (2013) found that two
decades of N additions (50 kg N/ha/yr) to a tundra meadow decreased both plant species
richness and plant community diversity.
6-130
DRAFT: Do Not Cite or Quote

-------
Table 6-18 Alpine and Arctic tundra plant diversity responses to nitrogen added via atmospheric deposition or
experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Arens et al. (2008)
Greenland
Dwarf-shrub/herb tundra
(Salix arctica, Carex
rupestris, Dryas integrifolia)
Addition
5, 10, 50
3
Plant community
composition
Not significant
Bassin et al. (2013)
Switzerland
Subalpine grassland
Addition
5, 10, 25, 50
7
Plant community
composition
Change
Armitaqe et al. (2014)
Europe (North
Atlantic)
Alpine heathlands
Ambient
0.6-39.6
n/a
Plant community
composition
Change
Sonq and Yu (2015)
China (Tibetan
Plateau)
Alpine meadow (Kobresia
humilis, Elymus nutans,
Stipa aliena, Festuca ovina)
Addition
3.75, 15, 75
8
Plant community
composition
Change
BishoD et al. (2010)
Washington (Mt. St.
Helens)
Primary successional alpine
meadow
Addition
78
5
Plant species
diversity
Increase
Bowman et al. (2012)
Colorado (Rocky
Mountain National
Park)
Dry sedge meadow
(Kobresia myosuroides,
Carex rupestris)
Addition
5, 10, 30
4
Plant species
diversity
Not significant
Bassin et al. (2013)
Switzerland
Subalpine grassland
Addition
5, 10, 25, 50
7
Plant species
diversity
Decrease
Sundavist et al. (2014)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Anthoxanthum alpinum)
Addition
100
3
Plant species
diversity
Not significant
Sundqvist et al. (2014)
Sweden
Tundra heath (Vaccinium
vitisidaea, Vaccinium
uliginosum, Betula nana)
Addition
100
3
Plant species
diversity
Not significant
June 2018
6-131
DRAFT: Do Not Cite or Quote

-------
Table 6 18 (Continued): Alpine and Arctic tundra plant diversity responses to nitrogen added via atmospheric
deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Farrer et al. (2015)
Colorado (Niwot
Ridge)
Moist alpine meadow
(Deschampsia cespitosa,
Geum rossii)
Addition
229
7
Plant species
diversity
Decrease
Wardleet al. (2013)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Empetrum hermaphroditum,
Vaccinium spp.)
Addition
50
21
Plant species
diversity
(vascular)
Decrease
Sona and Yu (2015)
China (Tibetan
Plateau)
Alpine meadow (Kobresia
humilis, Elymus nutans,
Stipa aliena, Festuca ovina)
Addition
3.75, 15, 75
8
Plant species
evenness
Not significant
Bowman et al. (2012)
Colorado (Rocky
Mountain National
Park)
Dry sedge meadow
(Kobresia myosuroides,
Carex rupestris)
Addition
5, 10, 30
4
Plant species
richness
Not significant
Armitaqe et al. (2014)
Europe (North
Atlantic)
Alpine heathlands
Ambient
0.6-39.6
n/a
Plant species
richness
Decrease
Sundavist et al. (2014)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Anthoxanthum alpinum)
Addition
100
3
Plant species
richness
Not significant
Sundavist et al. (2014)
Sweden
Tundra heath (Vaccinium
vitisidaea, Vaccinium
uliginosum, Betula nana)
Addition
100
3
Plant species
richness
Not significant
Sona and Yu (2015)
China (Tibetan
Plateau)
Alpine meadow (Kobresia
humilis, Elymus nutans,
Stipa aliena, Festuca ovina)
Addition
3.75, 15, 75
8
Plant species
richness
Low and mid
dose: not
significant
High dose:
decrease
June 2018
6-132
DRAFT: Do Not Cite or Quote

-------
Table 6 18 (Continued): Alpine and Arctic tundra plant diversity responses to nitrogen added via atmospheric
deposition or experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition Rate
(kg N/ha/yr)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Yuan etal. (2016)
Colorado (Niwot
Ridge)
Alpine meadow
Addition
50-200 (varied
over the 20 yr
duration;
averaged ca. 85)
20
Plant species
richness
Not significant in
dry and wet
meadow type;
decrease in
moist meadow
type
Southon etal. (2013) U.K.
Heathlands (Calluna
vulgaris)
Ambient
5.9-32.4
n/a
Plant species
richness
(vascular)
Decrease
Wardle et al. (2013) Sweden
Tundra meadow
(Deschampsia flexuosa,
Empetrum hermaphroditum,
Vaccinium spp.)
Addition
50
21
Plant species
richness
(vascular)
Decrease
ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-133
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
Because of the similarities in soil and plant community properties, heathlands are
discussed in this alpine and arctic tundra section. Southon et al. (2013) quantified plant
species richness at heathland sites in the U.K. along an N deposition gradient
(5.9-32.4 kg N/ha/yr) and observed declines in both vascular plant species richness and
bryophyte species richness. Similarly, Armitagc et al. (2014) observed changes in plant
community composition and decreases in plant species richness in alpine heathlands
along an N deposition gradient of 0.6 to 39.6 kg N/ha/yr across the northern U.K.,
Iceland, Norway, and the Faroe Islands. Three studies quantified changes in lichen
species richness in alpine tundra ecosystems along atmospheric N deposition gradients in
northern Europe; two observed significant declines with increases in N deposition. At 15
of the U.K. alpine tundra sites surveyed by Armitagc et al. (2014). Mitchell et al. (2016)
sampled soil microarthropods (oribatid, prostigmatid, and mesostigmatid mites, and
collembola). Across the gradient of sites, there was no direct influence of N deposition on
microarthropod community structure or the species richness, community composition, or
density for Collembola or any of the three groups of mites. However, N deposition
indirectly affected micrarthropod communities by altering variables such as graminoid
cover, moss depth, moss cover, and plant biomass C:N ratio, all of which significantly
influenced microarthropod communities (Mitchell et al.. 2016).
There has been a considerable amount of new research on the effects on added N plant
diversity in alpine meadows within the Tibetan Plateau region of China. In a
meta-analysis of N addition studies in this region, Fu and Shen (2016) found that N
additions decreased both plant species richness and plant diversity. Song and Yu (2015)
examined how different rates (3.75, 15, or 75 kg N/ha/yr) and chemical forms ofN
([NH4]2S04, NaNC>3, or NH4NO3) additions over 8 years influenced plant diversity and
the stability of plant communities in alpine plant communities on the Tibetan Plateau in
China. Throughout the experiment, there were no effects ofN form. The highest rate ofN
addition decreased community stability (mean biomass/mean temporal standard
deviation), species richness, and the dominance of community composition by individual
species; other rates ofN did not affect these metrics. Broadly, added N also increased the
temporal synchrony of species cover, meaning that there were stronger correlations
among species and functional groups in year-to-year variation in cover. This suggests that
N additions can decrease compensatory effects within plant communities, wherein one
species or functional group increases in cover during periods when others are exhibiting
decreased growth. Instead, the high N treatment increased the abundance of both of the
two dominant grass species. Song and Yu (2015) suggested that the reduction in
compensatory growth could result from decreased competition for N. As expected, there
was a strong correlation between log-transformed species variance in abundance and
log-transformed mean abundance, meaning that more abundant species were also more
variable in their abundance; this relationship was not affected by N additions. Notably,
6-134
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
there was no relationship between community stability and species richness. Although
this is unusual among studies of community stability and richness [e.g., (Lorcau and de
Mazancourt. 2013; Tilman et al.. 2006; Steiner et al.. 2005; Tilman. 1996)1. the high rate
of N addition caused the loss of only 2 of 20 species within these communities; neither
was a dominant species in the community. In comparison, 4 years of N additions
(100 kg N/ha/yr as NH4NO3) had no effect on plant species richness or diversity in a less
diverse (10-14 species) alpine meadow community on the Tibetan Plateau (Zona et al..
2016).
Among studies of microbial diversity (Table 6-19). three studies (Yuan et al.. 2016;
Farrer et al.. 2013; Nemergut et al.. 2008) investigated the effect of N additions in the
Rocky Mountains. Farrer et al. (2013) asked whether soil microbial changes were related
to the demonstrated decline in the abundance of the plant Geum rossi under increased N
deposition. Microbial community composition changed, but the influence of this change
on G. rossi was unclear. However, there was evidence that N additions imposed a
stronger physiological C limitation on G. rossi. Nemergut et al. (2008) also found shifts
in fungal and bacterial soil communities with chronic N additions (10 to 25 kg N/ha/yr
for >10 years). The fungal community shifted in response to N amendments, with a
decrease in the relative abundance of basidiomycetes. Bacterial community composition
also shifted in the N amended soil, with increases in the relative abundance of sequences
related to the Bacteroidetes and Gemmatimonadetes, and decreases in the relative
abundance of the Verrucomicrobia. Yuan et al. (2016) observed changes in bacterial
community composition after 20 years of N additions, with N additions increasing the
abundance of the phyla Chloroflexi and Bacteroidetes and N additions decreasing
Acidobacteria and Verrucomicrobia. Notably, forb biomass (but not grass biomass) was
positively correlated with bacterial diversity and bacterial richness covaried with plant
richness. Structural equation modeling revealed that the effects ofN additions on
bacterial community composition were primarily indirect effects occurring via changes in
soil pH and forb biomass rather than direct effects of N availability (Yuan et al.. 2016).
6-135
DRAFT: Do Not Cite or Quote

-------
Table 6-19 Alpine and Arctic tundra microbial diversity responses to nitrogen added via experimental
treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Pavne et al. (2012)
U.K.
Heathland (Calluna
vulgaris)
Addition
10, 20, 40,
80, 120
11, 21
Amoeba
community
composition
Change
Pavne et al. (2012)
U.K.
Heathland (Calluna
vulgaris)
Addition
10, 20, 40,
80, 120
11, 21
Amoeba species
diversity
Decrease
Pavne et al. (2012)
U.K.
Heathland (Calluna
vulgaris)
Addition
10, 20, 40,
80, 120
11, 21
Amoeba species
richness
Not significant
Pavne et al. (2012)
U.K.
Heathland (Calluna
vulgaris)
Addition
10, 20, 40,
80, 120
11, 21
Amoeba species
richness
Not significant
Yuan et al. (2016)
Colorado
(Niwot
Ridge)
Alpine meadow
Addition
50-200
(varied over
the 20 yr
duration;
averaged ca.
85)
20
Bacterial
phylotype
richness
Not significant in dry
and wet meadow type;
decrease in moist
meadow type
Nemerqut et al. (2008)
Colorado
(Niwot
Ridge)
Dry alpine meadow
(Kobresia myosuroides)
Addition
11.5
10
Fungal
community
composition
Change
Nemeraut et al. (2008)
Colorado
(Niwot
Ridge)
Dry alpine meadow
(Kobresia myosuroides)
Addition
11.5
10
Microbial
community
composition
Change
June 2018
6-136
DRAFT: Do Not Cite or Quote

-------
Table 6 19 (Continued): Alpine and Arctic tundra microbial diversity responses to nitrogen added via
experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition or
Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional
Nitrogen
Farreretal. (2013)
Colorado
(Niwot
Ridge)
Moist alpine meadow
(Deschampsia
cespitosa, Geum rossii)
Addition
288
11
Microbial
community
composition
Change
Wardleet al. (2013)
Sweden
Tundra meadow
(Deschampsia flexuosa,
Empetrum
hermaphroditum,
Vaccinium spp.)
Addition
50
21
Microbial
community
composition
Change
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-137
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
6.3.5.
Grasslands
In the 2008 ISA, there was consistent and coherent evidence suggesting that N additions
reduced plant biodiversity in grasslands in the U.S. and Europe. In the U.S., Clark and
Tilman (2008) evaluated the effects of chronic N addition over 23 years in Minnesota
prairie grasslands and found species numbers declined at the lowest addition level
(10 kg N/ha/yr added to 6 kg N/ha/yr of ambient deposition). In the San Francisco Bay
area of California, exotic nitrophilous grasses displaced native grass species, likely due to
greater N availability from deposition—at the time approximately 10 to
15 kg N/ha/yr—and from the cessation of grazing, which previously exported N out of
the ecosystem (Fcnn et al.. 2003b; Weiss. 1999).
In Europe, there were observations of N deposition-related declines in grassland plant
diversity in a variety of environments, including calcareous, neutral, and acidic soils,
species-rich heaths, and montane-subalpine grasslands (Stevens et al.. 2004; Bobbink et
al.. 1998; Bobbink et al.. 1992). In a transect of 68 acidic grasslands across Great Britain
that covered a broad range of ambient N deposition (5 to 35 kg N/ha/yr), chronic N
deposition significantly decreased plant species richness (Stevens et al.. 2004). Species
richness declined as a linear function of the rate of inorganic N deposition, with a
reduction of one species per 4-m2 quadrant for every 2.5 kg N/ha/yr of chronic N
deposition (Stevens et al.. 2004).
Although only a few studies had examined low levels of N inputs to grasslands in the
U.S. as of 2008, there were many that added higher levels of N (Clark et al.. 2007;
Bradley et al.. 2006; Suding et al.. 2005; Gough et al.. 2000) and examined how N
enrichment interacted with other factors such as fire, elevated atmospheric CO2, species
diversity, and climate change (Zavaleta et al.. 2003; Reich et al.. 2001; Collins et al..
1998). These studies were not necessarily designed to simulate N deposition, rather their
focus was on nutrient limitation and what happened following the alleviation of this
limitation. Recent gradient studies and studies using low level N addition rates verify that
the direction of effect between low and high N input rates are similar. This suggests that
although the magnitude of effect depends on the level of input, the direction of effect
does not, and that nutrient enrichment in grasslands follows a general response. This
response includes increases in aboveground production, decreases in light availability at
the soil surface, changes in soil fauna, and increases in litter buildup, all of which can
lead to competitive displacement of slower growing and shorter plant species from either
shading and/or reduced recruitment and regrowth (Hautier et al.. 2009; Bobbink et al..
1998; Tilman. 1993. 1987).
June 2018
6-138
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
In total, before 2008, large-scale biodiversity assessments across gradients of atmospheric
N deposition were restricted to Europe (Stevens et al.. 2004). Although there was wide
evidence in the U.S. linking the composition of plant and microbial communities to N
availability, most of these experiments had been conducted using N addition rates
considerably exceeding observed atmospheric N deposition rates in the U.S.
[e.g., (Bradley et al.. 2006; Suding et al.. 2005)1. Experiments demonstrating the
sensitivity of grassland community composition to lower rates of N deposition were more
limited, with evidence for Mediterranean grasslands (Weiss. 1999) and a northern prairie
ecosystem (Clark and Tilman. 2008).
Since the 2008 ISA, research on the effects of anthropogenic N on biodiversity in
grassland ecosystems has expanded to include a wider range of organisms and
ecosystems and a further understanding of how added N (particularly at more realistic N
input rates) alters biodiversity, including the species and functional groups most
susceptible to increased N availability. Much of this new research has been conducted in
the U.S. (particularly at Cedar Creek in Minnesota) and in western Europe, but China has
also emerged as a center for research on the effects of excess N on grasslands.
As noted earlier, Simkin et al. (2016) quantified the effect of N deposition on species
richness in the continental U.S. using data from over 15,000 plots, including grassland
ecosystems. While Europe has several large-scale continental studies on the effects from
N deposition (Stevens et al.. 2010a; Stevens et al.. 2004). this is the first study of its size
in the U.S. and provides a unique insight on the impacts from N deposition on U.S.
grassland ecosystems. In their analysis, Simkin et al. (2016) observed that at very low
rates of N deposition, increases in N deposition were associated with higher plant species
richness in open canopy ecosystems (e.g., grasslands, shrublands, and woodlands), but
species richness declined atN deposition rates above 8.7 kg N/ha/yr (5th-95th percentile:
6.4-11.3 kg N/ha/yr). Levels causing a decline in species richness are common across
much of the eastern and central U.S. and also occur near urban and intensive agricultural
areas in the West. Evidence of increases at low N input rates had not been observed
across N deposition gradients in Europe, likely due to higher historical deposition rates.
Simkin et al. (2016) also reported that the loss of species was strongly contingent on soil
pH, with more losses occurring in more acidic soils. Notably, the authors reported that
approximately 40% of grassland plots used for the study received N deposition at rates
above the threshold for declines in species richness.
A number of studies in Europe have built upon the work of Stevens et al. (2004) in
documenting N deposition impacts on grassland plant communities over long time
periods and/or large geographic areas, with these studies consistently indicating changes
in plant community composition and decreases in diversity as a result of N deposition
6-139
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
(see also Appendix 5 for discussion). Changes in grassland plant diversity in the U.K.
from N deposition have been particularly well documented. Stevens et al. (201 Ob) and
Maskell et al. (2010) built on the work of Stevens et al. (2004) in the U.K., trying to
identify whether changes in plant community composition caused by N deposition were
the result of acidification or eutrophication. Stevens et al. (2010b) worked from the same
68 acidic grasslands in Great Britain as Stevens et al. (2004). but included more detailed
information about soil pH and the sensitivity of individual plant species to pH and N
availability. Based on these data, Stevens et al. (2010b) observed that N deposition
mostly caused a shift toward more acid-tolerant species rather than nitrophilic species,
suggesting that acidification is the primary impact of N deposition on acidic grassland
plant communities. A broader analysis by Maskell et al. (2010). which expanded the
work of Stevens et al. (2004) and Stevens et al. (2010b') beyond acidic grasslands to also
include calcareous grasslands, heathlands, and mesotrophic grasslands—and also
included direct soil measurements at a subset of these sites—also found that the primary
link between N deposition and species loss was acidification in acidic grasslands and
heathlands. However, eutrophication was the primary link between species loss and N
deposition in calcareous grasslands (Maskell et al.. 2010). Notably, species losses along
the deposition gradient were smaller in calcareous ecosystems than in acidic ecosystems.
In a similar assessment, van Den Berg et al. (2016) observed negative effects of N
deposition on plant species richness in ecosystems across the U.K. (woodlands, heaths,
bogs, and grasslands), with the exception of a positive effect of deposition on species
richness in calcareous grasslands. Henrys et al. (2011) used data from two national
observation networks over Great Britain and found clear negative trends in plant species
prevalence to increasing N in all acidic grassland habitats. Field et al. (2014) also
expanded on the work of Stevens et al. (2004) by surveying a broader range of ecosystem
types across the U.K.: acidic grasslands, bogs, upland and lowland heaths, and sand
dunes. Higher N deposition was associated with decreases in species richness and
changes in community composition across all ecosystem types. Among functional
groups, N deposition decreased the diversity of mosses, lichens, forbs, and graminoids,
but generally increased the cover of graminoids (Field et al.. 2014). van Den Berg et al.
(2016) found that N deposition increased the ratio of grass-to-forb species in some U.K.
ecosystems (heathlands, bogs, acidic grasslands), but had the opposite effect in
calcareous grasslands.
Similar results were observed at broader scales in Europe. Several studies surveyed
153 acidic grassland sites across northern and western Europe (Pannek et al.. 2015;
Stevens et al.. 201 la; Stevens et al.. 201 lb. 2010a). These sites spanned a deposition
gradient of 2 to 44 kg N/ha/yr, but the decreases in plant species richness were not linear
and the highest rate of species loss occurred at deposition rates <20 kg N/ha/yr (Stevens
et al.. 2010a). Across the gradient, most of the decline in species richness was caused by
6-140
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
a loss of forb species, but grass and bryophyte species also declined (Stevens ct al..
2010a'). As a fraction of total species richness, grass species richness increased and forb
species richness declined with increasing N deposition (Stevens et al.. 2011b). Notably,
these changes occurred without consistent effects of N deposition on soil NOs . soil
NH4+, or tissue N concentrations in broadly sampled forb (Galium saxatile) and grass
(Agrostis capillaris) species, but both soil C:N and the foliar N concentration of a
bryophyte (Rhytidiadelphus squarrosus) were positively correlated with N deposition
(Stevens et al.. 2011b). Stevens et al. (201 lb) were able to explain 24% of the total
variation in plant species composition using climate, soil, and atmospheric deposition
metrics. Among this 24% of the total variation, soil variables (pH, aluminum
concentrations, C and N content) explained 38% of the variation and deposition
represented about 10% of this variation (Stevens et al.. 201 la). This influence of
deposition on community composition was notably weaker than the relationship to
changes in species richness. A survey of 44 species in these grasslands found that the
presence of 16 species responded to N deposition, with 12 of these 16 species responding
negatively (Pannek et al.. 2015). In the Atlantic region of France, Gaudnik et al. (2011)
observed that N deposition was one of the primary determinants of community
composition in acidic grasslands, even though the observed range of N deposition was
much smaller than in Stevens et al. (2004). Dupre et al. (2010) examined plant
community composition change in acidic grasslands in north-central Europe and the U.K.
over 70 years, dating back to 1939. Vegetation communities were differentiated
predominately according to soil pH, soil N availability, cumulative N deposition and S
deposition, and sampling date. Plot species richness declined through time for both
vascular plants and bryophytes, with cumulative N deposition identified as the primary
determinant of species loss.
While these N deposition gradient studies have provided strong documentation that
anthropogenic N pollution is altering biodiversity in grasslands in North America and in
Europe, experimental N addition studies have provided new information about the
mechanisms linking N deposition to shifts in community composition. For instance, using
grassland mesocosms in Switzerland receiving 150 to 200 kg N/ha/yr, Hautier et al.
(2009) conducted a unique experiment to understand the role of light competition in the
loss of species under conditions of high N availability. The addition of N increased
aboveground productivity and decreased both light availability near the ground surface
and species richness. An experimental increase in light availability caused by the use of
plant growth lights beneath the grassland canopy further increased productivity, but also
prevented the loss of plant species richness. Similarly, light competition was a
mechanism behind the negative effects of N deposition on species richness in grassland
sites in Israel (Demalach et al.. 2017).
6-141
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
At Cedar Creek in Minnesota, Reich (2009) conducted a 10-year experiment where
grassland assemblages of 16 perennial species were grown under factorial combinations
of ambient and elevated CO2 and ambient and elevated N. Increased N (40 kg N/ha/yr)
reduced species at both ambient and elevated CO2, by 16 and 8%, respectively. The
reduced losses of biodiversity at higher CO2 levels remains an active area of research but
is thought to be caused by greater soil water availability at higher CO2 levels (from
reduced stomatal conductance) that reduced the competitive displacement with N
addition. An even longer experiment, 25 years, conducted at Cedar Creek and analyzed
by Isbell et al. (2013a) also demonstrated that N enrichment decreased the number of
plant species, and that overtime, this effect became increasingly negative at all rates of N
addition. Moreover, species losses were nonrandom, with initially dominant native
perennial grasses becoming less dominant and then lost, and non-native perennial C3
grasses becoming increasingly dominant. Plot composition changed from a
high-diversity, native-dominated state (C4 grasses) to a low-diversity, non-native
dominated state (C3 grasses). Several earlier studies at this site also reported this general
effect (Clark and Tilman. 2008; Tilman. 1987). Nine native species were particularly
susceptible to becoming locally extinct under chronic nutrient enrichment, including
initially dominant and initially rare species. All of the findings from this seminal
experiment in Cedar Creek, MN, are representative of terrestrial eutrophication, as
opposed to acidification, because these soils were limed to maintain a constant pH
(Tilman. 1987). However, there were no lime additions in the experiment conducted at
this site by Reich (2009). indicating that both terrestrial eutrophication and acidification
could be occurring in that experiment.
Lan etal. (2015) examined how N deposition altered spatial patterns in biodiversity and
species-area relationships by measuring plant species loss and species richness in a
grassland in Inner Mongolia in 14 different plot sizes ranging from 1 to 25 m2. The
experiment included five levels of N addition between 17.5 and 280 kg N/ha/yr
(NH4NO3) for 10 years. Except for the 17.5 kg N/ha/yr treatment, N additions decreased
species richness. The absolute number of species lost versus the control group increased
rapidly as plot size increased from 1 to 8 m2, but then stayed the same (high N doses) or
decreased (low N doses). However, the proportional loss of species decreased as plot size
increased, which allowed for predictions of critical loads for species loss at different plots
sizes: 11.4 kg N/ha/yr at 5 m2 and 17.4 kg N/ha/yr at 25 m2. A previous study at these
sites using 0.5-m2 plots found a critical load of 8.5 kg N/ha/yr, indicating that species loss
as a consequence of N additions is sensitive to survey design.
Ecological theory suggests that plant diversity can be maintained by several factors,
including disturbance [e.g., (Davis et al.. 2000; Mack et al.. 2000; Hobbs and Huenneke.
1996)1. Among grasslands, the managed reintroduction of some disturbances has been
6-142
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
proposed as a way of maintaining plant diversity in eutrophic ecosystems. In a California
serpentine grassland, (Pasari et al.. 2014) examined the interactive effects of grazing and
simulated N deposition (80 kg N/ha total over 4 years) on native and exotic species
dynamics. With N additions, grazing helped maintain native species richness. High
grazing intensity decreased exotic plant cover, but only under ambient conditions (Pasari
et al.. 2014). In a southern California coastal grassland, Borer et al. (2014) conducted
several nutrient enrichment experiments, two of which involved the addition of N (40 and
100 kg N/ha/yr). At the higher N level, species richness (forbs primarily) was reduced by
one species per 0.5 m2 after 2 years. However, unlike in the work of Pasari et al. (2014).
this change in diversity was not affected by the presence of pocket gophers (Thomomys
bottae), a keystone herbivore that eliminated the effect of N on plant productivity. In an
annually burned Kansas tallgrass prairie, McLauchlan et al. (2014) analyzed a 27-year
record of plant community composition and N cycling. Despite rates of N deposition
(average 7 kg N/ha/yr) persistently within the range of critical loads for temperate
grasslands in the region, there was no evidence for increases in plant N concentrations,
decreases in forb diversity, or shifts in the relative abundance of dominant grass species.
The authors suggested that losses of N through frequent burning or grazing are sufficient
to prevent grassland eutrophication at low rates of N deposition.
There is also new evidence that N additions can alter flowering, seed production, and
seed abundance in grasslands. In Minnesota, Hillerislambers et al. (2009) developed a
statistical model to determine the effects of multiple global change factors, including N
deposition, on inflorescence mass. The effects of N deposition on seed production were
similar within functional groups and negatively correlated with aboveground
productivity: C3 grasses increased aboveground biomass and decreased seed production,
while C4 grasses decreased aboveground biomass and increased seed production. In a
greenhouse study of California oak savannah annual grasses, N additions (28 kg N/ha/yr
for 1 year) increased seed production (g/plant) by an invasive grass (Tulloss and
Cadenasso. 2016). In addition to seed production, the abundance of seed in soil seed
banks is important because seed banks help in revegetation following disturbances and
help maintain plant diversity, especially in small and isolated ecosystems (Piessens et al..
2004). However, N impacts on grassland seed banks are not necessarily directly related to
changes in plant cover or flowering. Basto et al. (2015a) found that 13 years of simulated
N deposition at rates of 35 or 140 kg N/ha/yr decreased cover and flowering in only one
species (Potentilla erecta) in an acidic grassland in England, but total seed bank
abundance declined by 34 and 61% in the low and high N treatments, respectively.
Further, seed bank species richness declined by 29 and 41% in these treatments. Among
taxa, there were decreases of >50% in the seed bank abundance of forbs, sedges, and
grasses with the high N dose, but only a significant decrease (34%) in grass seed bank
abundance at the lower N dose. Seed banks were quantified again 4 years after the
6-143
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
cessation of N additions, but did not significantly recover in any metric of abundance,
richness, or composition.
The Park Grass Experiment in England is the longest ecological experiment in the world
and has examined the effect of various soil amendments, including N (96 kg N/ha/yr) as
(NH4)2S04 or NaNC>3, on grassland productivity and composition since 1856. Zhalnina et
al. (2015) conducted 16S ribosomal RNA sequencing of soil samples from this
experiment to assess bacterial and archaeal community composition (Table 6-20). Soil
nitrate concentrations were positively correlated with the abundance of Thaumarchaeota
and Nitrospirae DNA, whereas total soil N was negatively correlated with Firmicutes,
Verrucomicrobia, and Chloroflexi. Multivariate analyses of community composition
found that the primary determinants of microbial community composition were soil pH
and C:N. Soil C:N was negatively correlated with Thaumarchaeota, Nitrospirae, and
Gemmatimonadetes, but positively correlated with Acidobacteria and Chlamydiae.
Changes in soil pH were a much larger control on microbial abundance, with significant
correlations with 17 of the 37 most abundant soil microbial genera. The two different
forms of N had different effects on pH, and unsurprisingly, different effects on microbial
community composition. Amendments of (NEL^SC^ decreased the phyla
Verrucomicrobia and Chloroflexi and the genera Bradyrhizobium, Paenibacillus, and
Geobacter, while NaNC>3 increased the abundance of the phyla Thaumarchaeota and
Nitrospirae and the genera Geobacter, Candidatus, Nitrososphaera, Nitrospira, and
Methylibium. In comparison, 27 years of N additions at Cedar Creek in Minnesota
changed bacterial composition (Fierer et al.. 2012; Ramirez et al.. 2010b') but had no
effect on bacterial diversity lYFierer et al.. 2012); Table 6-201.
6-144
DRAFT: Do Not Cite or Quote

-------
Table 6-20 Grassland microbial diversity responses to nitrogen added via experimental treatments.
Reference
Study Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional Nitrogen
Ramirez et al. (2010b)
Minnesota
(Cedar Creek)
Temperate
grassland
Addition
30, 60, 100,
160, 280,
500, 800
27
Bacterial
community
composition
Change
Fiereret al. (2012)
Minnesota
(Cedar Creek)
Temperate
grassland
Addition
34, 272
27
Bacterial
community
composition
Low dose: not sianificant
Hiah dose: chanae
Fiereret al. (2012)
Minnesota
(Cedar Creek)
Temperate
grassland
Addition
34, 272
27
Bacterial
diversity
Not significant
Wanqetal. (2017d)
China (Inner
Mongolia)
Steppe
grassland
Addition
50, 100, 150
8+
Ratio of fungal to
bacterial
abundances
Declined
Zhalnina et al. (2015)
U.K.
Temperate
grassland
Addition
96
153
Microbial
community
composition
Change
Daebeler et al. (2017)
Iceland
Subarctic
grassland
Addition
100
5
Microbial
community
composition
(archaeal
ammonia-
oxidizing
communities)
Not significant
June 2018
6-145
DRAFT: Do Not Cite or Quote

-------
Table 6 20 (Continued): Grassland microbial diversity responses to nitrogen added via experimental treatments.
Reference
Study Location Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional Nitrogen
Wanaetal. (2017d)
China (Inner
Mongolia)
Steppe
grassland
Addition
50, 100, 150
8+ Microbial
composition
(relative
abundance of
different
microbial groups)
Change; Actinomycetes,
Gram-negative bacteria, and
arbuscular mycorrhizae declined;
Gram-positive bacteria
increased; no significant change
in saprophytic fungi
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-146
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Relative to forests, there is comparably little information on how N additions shift the
composition of mycorrhizal communities in grasslands (Table 6-16). In a steppe
grassland in China, Chen et al. (2014) observed that 6 years of N additions
(100 kg N/ha/yr) decreased the richness of arbuscular mycorrhizal phylotypes found in
the soil, but did not affect the diversity of soil phylotypes or alter the richness or diversity
of phylotypes observed in plant roots.
As with that of forests, some information is also available regarding the effects of N
additions on grassland soil fauna (Table 6-17). In particular, several studies have been
conducted on soil fauna and their role in soil food webs at Cedar Creek in Minnesota
(Cesarz et al.. 2015; Eisenhauer et al.. 2013; Eisenhauer et al.. 2012). (Eisenhauer et al..
2012) observed that the N deposition treatment at Cedar Creek (40 kg N/ha/yr)
significantly altered the abundance of 5 of 14 groups of soil biota. Simulated N
deposition decreased predaceous nematode abundance by 62%, but increased fungal
feeding nematodes by 206%. Nematode taxon richness and microarthropod taxon
richness declined by 7 and 15% under added N (Eisenhauer et al.. 2012). A more detailed
analysis by Cesarz et al. (2015) found that N enrichment increased the density of the
plant-feeding Longidoridae nematode family by 148% and increased the density of a
rapidly growing guild of fungal-feeding nematodes, while a slower-growing guild of
predaceous nematodes declined (Cesarz et al.. 2015). Overall, N additions increased
nematode community composition toward nematode guilds favored by decomposition
pathways dominated by fungi and away from bacterial-dominated decomposition
pathways (Cesarz et al.. 2015; Eisenhauer et al.. 2012). However, while simulated N
deposition had significant effects on soil food webs, these influences were weaker than
the direct effects caused by differences in plant diversity (Eisenhauer et al.. 2013).
6.3.6. Arid and Semiarid Ecosystems
As of the 2008 ISA, there was consistent and coherent evidence of altered plant
communities in N addition experiments in arid and semiarid ecosystems, particularly
within CSS and chaparral ecosystems along the southern California coast and in portions
of the Mojave Desert near large population centers. There was further additional evidence
about shifts in plant and microbial community composition from N addition experiments
in the Mojave, Chihuahuan, Sonoran, and Great Basin Deserts. Although many of these
ecosystems are relatively remote and lightly impacted by anthropogenic N deposition, the
2008 ISA noted that deposition to some arid and semiarid ecosystems can be high
downwind of major urban and agricultural areas, reaching more than 30 kg N/ha/yr in
areas of southern California (Fenn et al.. 2003b). Like other terrestrial ecosystems, there
was widespread evidence that N additions altered plant communities by causing a
June 2018
6-147
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
differential stimulation of growth among plant species (Bacz et al.. 2007; Inouve. 2006).
such as by favoring rapidly growing nitrophilous species (Fenn et al.. 2003b'). In addition
to these effects on plant communities, Egerton-Warburton and Allen (2000) found a shift
in arbuscular mycorrhizal community composition with decreased species richness and
diversity along an N deposition gradient among CSS ecosystems in southern California (2
to 57 |ig N/g as soil NOs ). These shifts in mycorrhizal fungal communities may facilitate
replacement of native plant communities by exotic annual grasslands.
There are two ecological interactions that are particularly important controlling how N
deposition influences plant community composition in arid and semiarid ecosystems:
(1)	the strong dependence of biological responses on an adequate water supply and
(2)	the ability of N deposition to promote the growth of exotic plants [particularly annual
grasses; (Brooks. 2003)1. which can dramatically alter the fire cycle by providing a more
spatially continuous fuel supply (Brooks et al.. 2004). The fire cycle impacts of increased
invasive grass biomass were particularly apparent in southern California in CSS
ecosystems near the coast and in Mojave Desert ecosystems inland (Brooks et al.. 2004;
Brooks and Esque. 2002; Cione et al.. 2002; Yoshida and Allen. 2001; Brooks. 1999;
Eliason and Allen. 1997). Fire was relatively rare in the Mojave Desert until the past two
decades, but fire now occurs frequently in areas that have experienced invasion of exotic
grasses (Brooks. 1999). These interactions are apparent in an NH4NO3 addition
(32 kg N/ha/yr) study in the Mojave Desert of southern California conducted by Brooks
(2003). In the second and wetter year of the experiment, N additions decreased species
richness among native annual plants (Brooks. 2003). Overall, N additions increased the
growth of the invasive grass compact brome (Bromus madritensis) beneath the dominant
native shrub creosote bush (Larrea tridentata) and increased the growth of invasive
grasses in the genus Schismus and the invasive forb Erodium cicutarium in the
interspaces between Larrea shrubs, creating a more continuous fuel bed.
Since the 2008 ISA, a number of new N addition experiments and ambient N deposition
gradient studies on shifts in plant community composition and diversity have been
conducted, primarily in California and in arid portions of China (Table 6-21). Broadly,
while some of these studies found decreases in plant species richness [e.g., (Sun etal..
2014; Allen et al.. 2009)1. there were more frequent observations of changes in plant
community composition. This disparity may be a function of both the relative brevity of
these experiments (mostly 2 to 4 years) and the strong moisture limitation that constrains
biological responses to added N in these ecosystems.
6-148
DRAFT: Do Not Cite or Quote

-------
Table 6-21 Arid and semiarid ecosystem plant diversity responses to nitrogen added via atmospheric
deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional Nitrogen
Rao et al. (2009)
California
(Joshua Tree
NP)
Creosote bush (Larrea
tridentata) or pinyon-juniper
woodland (Pinus monophytta,
Juniperus californica)
Ambient
2.7-14.4
n/a
Invasive annual
grass cover
Increase
Allen et al. (2009)
California
(Joshua Tree
NP; four sites)
Creosote Bush (Larrea
tridentata) scrub; pinyon-
juniper woodland (Pinus
monophyiia, Juniperus
californica)
Addition
5, 30
2
Native plant
species richness
Low dose: not sianificant
Hiqh dose: not sianificant at
two sites, increase at one site,
decrease at one site
Vourlitis (2017)
California
(southern,
coastal)
Coastal sage scrub
(.Artemisia californica, Salvia
mellifera)
Addition
50
13
Native and
exotic plant
species cover
Increase in the native shrub
Artemesia californica in the 4th
and 5—9th yr of the 13-yr
experiment; decrease in the
native shrub Salvia mellifera in
the 4th and 11-13th yr;
increase in the exotic plant
Brassica nigra in the 11 -13th
yr
Vourlitis and
Pasquini (2009)
California
(southern
coastal)
Coastal sage scrub
(.Artemisia californica, Salvia
mellifera)
Addition
50
5
Plant community
composition
Change
Vourlitis and
Pasauini (2009)
California
(southern
coastal)
Chaparral (Adenostoma
fasciculatum, Ceanothus
greggii)
Addition
50
5
Plant community
composition
Not significant
June 2018
6-149
DRAFT: Do Not Cite or Quote

-------
Table 6 21 (Continued): Arid and semiarid ecosystem plant diversity responses to nitrogen added via
atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional Nitrogen
Pasauini and
Vourlitis (2010)
California
(southern;
three sites)
Chaparral (Adenostoma
fasciculatum, Ceanothus
spp.)
Ambient
8.1, 11.9,
18.4
n/a
Plant community
composition
Change
Concilio and Loik
(2013)
California
(Great Basin)
Cheatgrass (Bromus
tectorum) in sagebrush
(.Artemesia tridentata) steppe
Addition
50
4
Plant community
composition
Not significant
Cox etal. (2014)
California
(southern
coastal)
Coastal sage scrub
Ambient
5.7-23.8
n/a
Plant community
composition
More invasive grasses when N
deposition >11 kg N/ha/yr
Ochoa-Hueso and
Stevens (2015)
Spain
Shrubland (Quercus
coccifera, Rosmarinus
officinalis, Lithodora
fruticosa)
Addition
10, 20, 50
3
Plant community
composition
Change
Zhana etal. (2015b)
China (north,
Songnen)
Alkaline grassland (Leymus
chinensis, Kalimeris
integrifolia)
Addition
100
4
Plant community
composition
Change
Collins etal. (2017)
New Mexico
Grassland (Bouteloua
eriopoda, Bouteloua gracilis)
Addition
20
1 to 7
Plant community
composition
Change in 3 of the 4 yr
following fire; not significant in
2 yr preceding fire and last
year of the experiment
McHuqh et al.
(2017)
Utah
Semiarid grassland
Addition
2, 5, 8
2
Plant community
composition
Not significant
Zhana etal. (2015b)
China (north,
Songnen)
Alkaline grassland (Leymus
chinensis, Kalimeris
integrifolia)
Addition
100
4
Plant community
evenness
Not significant
Zhana etal. (2014)
China (north,
Inner
Mongolia)
Alkaline grassland (Leymus
chinensis, Stipa grandis)
Addition (2
or 12
additions/yr)
10, 20, 30,
50, 100, 150,
200, 500
5
Plant species
diversity
Decrease (stronger decrease
with two additions/yr)
June 2018
6-150
DRAFT: Do Not Cite or Quote

-------
Table 6 21 (Continued): Arid and semiarid ecosystem plant diversity responses to nitrogen added via
atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional Nitrogen
Ochoa-Hueso and
Stevens (2015)
Spain
Shrubland (Quercus
coccifera, Rosmarinus
officinalis, Lithodora
fruticosa)
Addition
10, 20, 50
3
Plant species
diversity
Not significant
Zhanq et al. (2015b)
China (north,
Songnen)
Alkaline grassland (Leymus
chinensis, Kalimeris
integrifolia)
Addition
100
4
Plant species
diversity
Decrease
Vourlitis and
Pasauini (2009)
California
(southern
coastal)
Coastal sage scrub
(Artemisia caiifornica, Salvia
mellifera)
Addition
50
5
Plant species
richness
Not significant
Vourlitis and
Pasauini (2009)
California
(southern
coastal)
Chaparral (Adenostoma
fasciculatum, Ceanothus
greggii)
Addition
50
5
Plant species
richness
Not significant
Concilio and Loik
(2013)
California
(Great Basin)
Cheatgrass (Bromus
tectorum) in sagebrush
(Artemesia tridentata) steppe
Addition
50
4
Plant species
richness
Not significant
Sun et al. (2014)
China (north,
Songnen)
Shrubland (Leymus
chinensis, Artemisia
scoparia)
Addition
23, 46, 69, 92
3
Plant species
richness
Low dose: not sianificant
Other doses: decrease
Zhanq et al. (2014)
China (north,
Inner
Mongolia)
Alkaline grassland (Leymus
chinensis, Stipa grandis)
Addition (2
or 12
additions/yr)
10, 20, 30,
50, 100, 150,
200, 500
5
Plant species
richness
Decrease (stronger decrease
with two additions/yr)
Zhana et al. (2015b)
China (north,
Songnen)
Alkaline grassland (Leymus
chinensis, Kalimeris
integrifolia)
Addition
100
4
Plant species
richness
Decrease
McHuah et al.
(2017)
Utah
Semiarid grassland
Addition
2, 5, 8
2
Plant species
richness
Not significant
June 2018
6-151
DRAFT: Do Not Cite or Quote

-------
Table 6 21 (Continued): Arid and semiarid ecosystem plant diversity responses to nitrogen added via
atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or Addition
Nitrogen
Addition
Rate (kg
N/ha/yr)
Duration
(yr)
Endpoint
Effect of Additional Nitrogen
Vourlitis (2017)
California
Coastal sage scrub
Addition
50
13
Plant species
Decrease in the last 3 yr of the

(southern,
(.Artemisia californica, Salvia



richness
13-yr experiment; not

coastal)
mellifera)




significant for previous years
Collins etal. (2017)
New Mexico
Grassland (Bouteloua
eriopoda, Bouteloua gracilis)
Addition
20
1 to 7
Plant species
richness for
grasses and
forbs
Not significant
ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; NP = national park; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only statistically significant effects are listed as increases
or decreases.
June 2018
6-152
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Within southern California, there was a continued research focus on how N deposition
alters the abundance of invasive annual plants (particularly grasses) in CSS, chaparral,
and Mojave Desert ecosystems (Cox et al.. 2014; Pasquini and Vourlitis. 2010; Allen et
al.. 2009; Rao et al.. 2009; Vourlitis and Pasquini. 2009; Talluto and Suding. 2008).
Among inland ecosystems, Allen et al. (2009) added N (5 and 30 kg N/ha/yr for 2 years)
to four sites at Joshua Tree National Park (JOTR) in California and found that the effects
of added N on native plant diversity depended on the abundance of invasive grasses. At
one site where invasive grass biomass (primarily Bromus madritensis) was high and
responded positively to added N, N additions decreased native forb species richness; at a
site where invasive grasses were a small component of the plant community, N additions
increased native forb cover and species richness. Given that Rao et al. (2009) found
greater invasive annual grass cover at sites with higher N deposition across a 16-site N
deposition gradient in the JOTR area, the results of Allen et al. (2009) imply that N
deposition may be decreasing native forb richness across broad areas in the JOTR area.
Near the southern California coast, Vourlitis and Pasquini (2009) investigated the effects
of dry-season N addition in chaparral (dominated by the shrubs Adenostoma fasciculatum
and Ceanothus greggii) and CSS (dominated by the shrubs Artemisia californica and
Salvia mellifera) stands over a 5-year period. The additional N (50 kg N/ha/yr)
significantly altered the community composition of CSS, but not chaparral. The
differences in CSS community composition were due to changes in the relative
abundance of dominant shrubs, not herbaceous plant species. This is not necessarily
consistent with previous research, including observations that although both native shrubs
and invasive annual plants can exhibit faster growth rates with added N, high N
availability can cause mortality in native shrubs (Allen etal.. 1998). Talluto and Suding
(2008) revisited 232 plots in southern California that had been dominated by CSS
vegetation during a 1930s vegetation survey and found that CSS cover had declined by
49%, and only 15% of the 1930s CSS plots lacked invasive annual grasses. At a
landscape scale, the conversion of CSS to invasive-dominated land was positively related
to N deposition rates and fire frequency, which itself has been linked to N deposition. In
another landscape-level analysis of vegetation change since the 1930s, Cox et al. (2014)
observed that CSS converted to non-native grasslands when N deposition was greater
than 11 kg N/ha/yr, more consistent with the previous results. Also in southern
California, Pasquini and Vourlitis (2010) found evidence that N deposition caused
changes in chaparral plant community composition in recently burned ecosystems, but
the influence of N deposition in this study was confounded by changes in other
environmental factors across the three-site N deposition gradient (8.1 to 18.4 kg N/ha/yr)
used for this research.
6-153
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Elsewhere in California, Concilio and Loik (2013) studied the effects of added N
(50 kg N/ha/yr for 4 years) on sagebrush steppe ecosystems in the eastern Sierra Nevada
where the invasive annual grass Bromus tectorum was beginning to establish. Although N
addition treatments increased plant available N and total N, no changes were observed in
cover, community composition, or richness of native or invasive species. Bromus
tectorum cover was inversely related to native forb species richness, but increased N
deposition did not affect plant diversity. However, Concilio and Loik (2013) cautioned
that precipitation was below average during the study and N additions may have different
effects when moisture availability is greater.
Several notable N addition studies have been conducted in the arid steppe grassland and
shrubland ecosystems in northern and western China. Zhang et al. (2014) studied whether
the frequency of additions (12 doses/year vs. 2 doses/year) altered the effect of added N
on plant diversity in a steppe grassland in the Inner Mongolia region of China (eight
levels from 10 to 500 kg N/ha/yr as NH4NO3 for 5 years). Species richness declined by
about 50% as N addition rates increased; plant diversity also decreased more with higher
N addition rates. Zhang et al. (2014) noted that the declines in diversity and richness were
smaller with the smaller frequent N doses than with the large infrequent doses, and
suggested that this implied that some N addition studies may be overestimating the
negative effects on plant biodiversity. However, at the lower N addition rates (10, 20, and
30 kg N/ha/yr) that are more relevant to N deposition levels in the U.S., the effects of
dose frequency were not consistently observed; species richness was sometimes
(nonsignificantly) higher with the large infrequent doses at these N addition rates. The
frequency of N additions did not impact the rate of species loss in perennial forbs, but did
affect the rate of species loss in grasses and annual and biennial forbs. Notably, N
additions were associated with significant decreases in soil temperature, soil moisture,
and soil pH and increases in NH4+ and NO3 . All of these soil changes were correlated
with each other and the proportional changes in each of these variables were correlated
with the proportional loss in species richness. Among these soil traits, higher N addition
frequency increased NH44", but had no other significant effects. In subsequent research at
these sites, Zhang et al. (2016b) observed that the lower plant species richness in the low
frequency N addition plots relative to the high frequency plots was not caused by
differences in the rate of species loss, but instead resulted from slower colonization by
new species in the plots that received large infrequent N additions. In another experiment
in semiarid Chinese steppe grasslands, Li et al. (2016a) observed thatN additions (50,
100, 150 kg N/ha/yr for 8 years) decreased species richness, decreased diversity, and
altered community composition for both plant and bacterial communities and that
changes in plant and microbial communities were significantly correlated. Li et al.
(2016a) also observed that N additions decreased both plant and microbial species
6-154
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
richness and altered microbial community composition in a semiarid Chinese steppe
grassland.
A recent study by Tian et al. (2016b) examined how N addition at several rates affected
grassland structure and function in the temperate steppe of China (20, 40, 80, 160,
320 kg N/ha/yr). They found that N addition led to increases in aboveground biomass and
to shifts in relative abundance from forbs to grasses even at the lowest input rate, as
reported in other studies. However, they reported a novel mechanism related to
manganese (Mn) toxicity to explain reductions in forb diversity that had not been
demonstrated before, but had been reported in forestry literature (St.Clair et al.. 2008;
St.Clair and Lynch. 2005). They reported that forbs were much more prone to Mn
toxicity than grasses following N addition because of physiological differences in root
enzymes associated with metal uptake. Manganese accumulated more in forbs than
grasses, leading to decreased forb photosynthetic rates and shifts in relative abundances
towards grasses. It is unknown whether this mechanism is operating in other grassland
systems.
In an arid steppe grassland in northeastern China, Sun et al. (2014) observed that N
additions of 23, 46, 69, and 92 kg N/ha/yr (3 years, as urea) caused proportional
decreases in forb species richness, but found added N caused soil bacterial communities
to become more diverse except at the highest N dose (Table 6-22). Huang et al. (2015)
conducted a 3-year N x water experiment focused on soil microbial community
composition in a desert steppe ecosystem in northwestern China that receives
25 kg N/ha/yr of ambient deposition. The N addition treatment (50 kg N/ha/yr as
NH4NO3) effects on microbial composition differed between inter-plant and beneath
shrub sites. In the interspaces, N additions increased the portion of bacteria in total
microbial biomass in all 3 years and increased actinobacteria in the last 2 years of the
experiment. Beneath the shrubs, N additions broadly suppressed microorganisms in all
domains. In the Sonoran Desert near Phoenix, Marusenko et al. (2015) found that
although N additions (60 kg N/ha/yr for 8 years) increased the abundance of the amoA
gene (needed for ammonia oxidation) in both archaea and bacteria, the community
composition of ammonia-oxidizing microorganisms was unaffected.
In Mediterranean shrublands in Spain, N additions (10, 20, and 50 kg N/ha/yr for 4 years)
increased the abundance of Collembola, which were a dominant component (44%) of the
soil fauna. The N additions did not significantly affect soil fauna richness or diversity, but
there was a significant negative relationship between richness and soil C:N ratio and a
positive relationship between diversity and soil pH (Ochoa-Hueso et al.. 2014).
6-155
DRAFT: Do Not Cite or Quote

-------
Table 6-22 Arid and semiarid ecosystem microbial diversity responses to
nitrogen added via experimental treatments.


Ambient
Nitrogen




Deposition
Addition

Effect of

Study
or
Rate (kg
Duration
Additional
Reference
Location
Vegetation Addition
N/ha/yr)
(yr)
Endpoint Nitrogen
Marusenko et al.
(2015)
Phoenix,
AZ
Creosote and
bursage
shrublands
(Larrea
tridentata,
Ambrosia spp.
Addition
60
Archaeal
and bacterial
amoA (NH3
mono-
oxygenase)
community
composition
Not
significant
Marusenko et al. Phoenix, Creosote and Addition
(2015)	AZ	bursage
shrublands
(Larrea
tridentata,
Ambrosia spp.)
60
Archaeal
and bacterial
amoA (NH3
mono-
oxygenase)
gene
abundance
Increase
Huana et al. (2015)
China
Desert shrubs
(Haloxylon
ammodendron)
Addition
50

3
Microbial
community
composition
Change
McHuqh et al.
(2017)
Utah
Semiarid
grassland
Addition
2, 5, 8

2
Microbial
community
composition
Not
significant
McHuqh et al.
(2017)
Utah
Semiarid
grassland
Addition
2, 5, 8

2
Microbial
diversity
Not
significant
Sun et al. (2014)
China
(north,
Songnen)
Shrubland
(Leymus
chin en sis,
Artemisia
scoparia)
Addition
23, 46, 69,
92
3
Soil bacterial
community
composition
Change
Sun et al. (2014)
China
(north,
Songnen)
Shrubland
(Leymus
chin en sis,
Artemisia
scoparia)
Addition
23, 46, 69,
92
3
Soil bacterial
diversity
Increase
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only
statistically significant effects are listed as increases or decreases.
June 2018
6-156
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
6.3.7. Lichens
In the 2008 ISA, there was consistent and coherent evidence indicating that lichen
communities were affected by current levels of N deposition. Lichen community
composition is highly sensitive to atmospheric pollution from N and S (Jovan. 2008).
Based on the widely recognized sensitivity of lichen species to air pollution, Geiser and
Neitlich (2007) developed a model to assess local air quality in Pacific Northwest forests
based on lichen community composition. In addition to being good subjects for
biomonitoring, lichens constitute important components of the forest ecosystem by
contributing to biodiversity, regulating nutrient and hydrological cycles, and providing
habitat elements for wildlife (McCunc and Geiser. 1997). The composition of the lichen
community is important because individual species have different physical and
physiological traits, and thus make particular contributions to the provisioning of
ecosystem services.
Lichens containing a cyanobacterial photobiont appear to be more sensitive to adverse
effects from atmospheric N deposition than most other lichens (Hallingback and Kellner.
1992; Hallingback. 1991). In central Europe, Hauck and Wirth (2010) analyzed
514 lichen species and found that shade-adapted lichen species are nearly universally
intolerant of high rates of N deposition. Nearly all of these shade-tolerant and
pollution-sensitive species are crustose lichens (Hauck and Wirth. 2010). The decline of
lichens containing cyanobacteria in parts of northern Europe has been associated with N
deposition in the range of 5 to 10 kg N/ha/yr (Bobbink et al.. 1998). In Sweden, the
probability of occurrence for three genera of hair lichens (Alectoria, Bryoria, Usnea) on
Norway spruce (Picea abies) in forest inventory plots peaked at 3-6 kg N/ha/yr of
deposition and then rapidly declined; N deposition was the strongest or second-strongest
environmental predictor of lichen occurrence (Esseen et al.. 2016). In the U.S., lichen
species are negatively affected by N inputs as low as 3 to 8 kg N/ha/yr (Fenn et al..
2003a).
In the San Bernardino Mountains, CA, up to 50% of lichen species that occurred in the
region in the early 1900s have disappeared (Fenn et al.. 2003a; Nash and Sigal. 1999). In
mixed conifer forests in California, the critical load for lichen communities has been
estimated at 3.1 kg N/ha/yr (Fenn et al.. 2008). Compared to California, air pollution has
historically been less problematic in the Pacific Northwest and this region still has large
populations of pollution-sensitive lichens (Jovan. 2008; Fenn et al.. 2003a). Jovan (2008)
reported that hotspots for lichen diversity in the U.S. include Klamath-Siskiyou region
along the Oregon-California border, the Okanogan highlands region in northeastern
Washington, and the Blue Mountains in eastern Oregon—all areas that are relatively
distant from large human population centers and industrial centers. However, lichen
June 2018
6-157
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
communities in the Pacific Northwest are beginning to show evidence of changes in
response to increased N pollution, including decreased distribution of sensitive lichen
taxa and their replacement with nitrophilous species (Geiser and Neitlich. 2007).
Research on lichens since 2008 continues to indicate that lichen communities change in
response to increased atmospheric N, in both U.S. and Europe. Unlike most other N
deposition biodiversity studies, lichen research has been dominated by studies using
measurements along ambient N deposition gradients rather than using experimental N
additions (Table 6-23). Within the U.S., there are numerous examples of shifts in lichen
community composition along gradients of atmospheric N deposition. Lichen community
composition has been quantified in many parts of the U.S. as part of the U.S. Department
of Agriculture Forest Service's Forest Inventory and Analysis (FIA) program. Jo van
(2008) reported the results from surveys of almost 800 plots conducted from 1998 to
2003 in California, Washington, and Oregon, documenting clear shifts in lichen
community composition in forests in and around areas of intensive agricultural and
industrial production in the Central Valley of California, the Willamette Valley in
Oregon, and the Puget Trough (Seattle/Tacoma/Olympia) in Washington. In the
northeastern U.S., Will-Wolf et al. (2015) analyzed approximately 600 plots of survey
data collected in the 1990s and 2000s as part of the FIA program and from other sources.
These data showed strong relationships between total N deposition and decreased lichen
abundance, reduced species richness, and altered community composition, but the high
overlap between N deposition and acidifying deposition in this region made it difficult to
discern the primary influence on lichen biodiversity.
Table 6-23 Lichen biodiversity responses to nitrogen added via atmospheric
deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr or
as Otherwise
Noted)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Jovan (2008)
California,
Oregon,
Washington
Forest
Ambient
0.5-21
n/a
Community
composition
Change
Roqers et al.
(2009)
Utah and
Idaho
Aspen forests
(Populus
tremuloides)
Ambient
(NH3 gas)
7.3-92.2 pm/m3
n/a
Community
composition
Change
Geiser et al.
(2010)
Oregon and
Washington
Conifer forests
Ambient
0.8-8
n/a
Community
composition
Change
June 2018
6-158
DRAFT: Do Not Cite or Quote

-------
Table 6-23 (Continued): Lichen biodiversity responses to nitrogen added via
atmospheric deposition or experimental treatments.




Nitrogen






Ambient
Addition Rate






Deposition
(kg N/ha/yr or


Effect of

Study

or
as Otherwise
Duration

Additional
Reference
Location
Vegetation
Addition
Noted)
(yr)
Endpoint
Nitrogen
Johansson et
Sweden
Norway spruce
Addition
6, 12.5, 25, 50
4
Community
Change
al. (2012)

(Picea abies)



composition

Jovan et al.
California
Oak forests
Ambient
6.1-71.1
n/a
Community
Change
(2012)
(Los
(Quercus



composition


Angeles
kelloggii)






Basin)






Gibson et al.
Nova
Northern
Ambient
0.01-0.35 ppb
n/a
Community
Change
(2013)
Scotia
hardwood
(NO2 gas)


composition


(Canada)
forests (Acer







saccharum,







Betula







alleghaniensis)





Schirokauer et Alaska Conifer forests Ambient	0.05-1.05	n/a Community Change
al. (2014b) (southeast)	composition
McMurrav et Idaho, Conifer forests Ambient	0.5-4.3	n/a Community Change
al. (2015) Wyoming,	composition
Montana
Johansson et Sweden Norway spruce Addition 6,12.5,25,50 4 Species 6 and
al. (2012)	(Picea abies)	richness 12.5 doses:
increase
25 and
50 doses:
decrease
Gibson et al.
(2013)
Nova
Scotia
(Canada)
Northern
hardwood
forests (Acer
saccharum,
Betula
alleghaniensis)
Ambient
(NO2 gas)
0.01-0.35 ppb
n/a
Species
richness
Decrease
Southon et al.
(2013)
U.K.
Heathlands
(Calluna
vulgaris)
Ambient
5.9-32.4
n/a
Species
richness
Decrease
Armitaqe et al.
(2014)
Europe
(North
Atlantic)
Alpine
heathlands
Ambient
0.6-39.6
n/a
Species
richness
Not
significant
Field et al.
(2014)
U.K.
Heathlands
Ambient
5.4-32.4
n/a
Species
richness
Decrease
McDonouqh
and
Watmouah
(2015)
Ontario,
Canada
Sugar maple
forests (Acer
saccharum)
Ambient
8.3-12.9
n/a
Species
richness
Not
significant
June 2018
6-159
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
Table 6-23 (Continued): Lichen biodiversity responses to nitrogen added via
atmospheric deposition or experimental treatments.
Reference
Study
Location
Vegetation
Ambient
Deposition
or
Addition
Nitrogen
Addition Rate
(kg N/ha/yr or
as Otherwise
Noted)
Duration
(yr)
Endpoint
Effect of
Additional
Nitrogen
Will-Wolf et al.
(2015)
North-
eastern
U.S.
Forests
Ambient
Not stated
n/a
Species
richness
Decrease
ha = hectare; kg = kilogram; N = nitrogen; n/a = not applicable; NH3 = ammonia; N02 = nitrogen dioxide; ppb = parts per billion;
yr = year.
Notes: single studies are reported more than once if multiple endpoints were measured. References ordered by endpoint. Only
statistically significant effects are listed as increases or decreases.
Other surveys of lichen community composition in the U.S. also documented changes
even at relatively low rates of N deposition. For instance, lichen community composition
in southeastern Alaska shifted toward more eutrophic species along an N deposition
gradient of 0.05 to 1.05 kg N/ha/yr (Schirokaucr et al.. 2014b). In the northern Rockies,
McMurrav et al. (2015) observed a shift in lichen community composition at deposition
rates greater than 4.0 kg N/ha/yr. In montane aspen forests in the Utah-Idaho border
region, Rogers et al. (2009) observed a shift in lichen community composition away from
nitrophilous species along a NH3 concentration gradient away from urban and
agricultural sources. In the Los Angeles Basin area of California, Jovan et al. (2012)
documented increases in the abundance of eutrophic lichen species along a gradient of N
deposition. In coniferous forests in the Pacific Northwest, shifts in lichen community
composition and declines in oligotrophic lichen species were observed as N deposition
increased to levels from 3 to 9 kg N/ha/yr range (Geiser et al.. 2010; Glavich and Geiser.
2008).
Outside of the U.S., changes in lichen community composition have been observed in
Canada and in Europe. In the U.K., Southon et al. (2013) and Field et al. (2014) observed
decreased lichen species richness in over 50 heathland survey locations along an N
deposition gradient of 5.4 to 32.4 kg N/ha/yr. However, Armitage etal. (2014) did not
find a significant change in lichen species richness along an N deposition gradient of 0.6
to 39.4 kg N/ha/yr in alpine heathlands in the North Atlantic region of Europe. In
Sweden, Johansson et al. (2012) found additions of 6 kg N/ha/yr, applied directly onto
trees in a liquid spray, in old growth boreal spruce forests changed the species
composition of epiphytic lichen communities and reduced species richness. Gibson et al.
(2013) also noted that in Cape Breton Highlands National Park in Nova Scotia, the
number of pollutant-intolerant lichen species decreased from 10 to 5 when summer and
winter atmospheric NO2 concentrations increased to 0.46 and 0.15 ppb, respectively.
McDonough and Watmough (2015) were unable to detect an influence of N deposition
June 2018
6-160
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
on epiphytic foliose lichen species richness across a network of 40 sugar maple forest
monitoring plots across Ontario, Canada. However, study sites in areas with a history of
intense industrial activity tended to have very low epiphytic foliose lichen species
richness.
6.3.8. Most Sensitive Ecosystems
The 2008 ISA reported that the most responsive ecosystems to N enrichment from
atmospheric N deposition were those that receive high levels of N loading, are N limited,
or contain species that have evolved in nutrient-poor environments. Species adapted to
low N supply are readily outcompeted by species that have higher N demand when the
availability of N is increased (Krupa. 2003; Tilman and Wedin. 1991; Aerts et al.. 1990).
As a consequence, some native species can be eliminated by N deposition (Stevens et al..
2004; Falkengren-Grerup. 1989. 1986; Roelofs. 1986; Ellenberg. 1985).
Unlike the case of acidifying deposition where ecosystem sensitivity is tied principally to
underlying geology, most terrestrial ecosystems are N limited and, therefore, sensitive to
perturbation caused by N additions (Le Bauer and Treseder. 2008). Consequently, little
was known in the 2008 ISA about the full extent and distribution of the terrestrial
ecosystems in the U.S. most sensitive to adverse impacts caused by atmospheric N
deposition. Effects were most likely to occur where areas of relatively high atmospheric
N deposition intersect with N limited plant communities. The factors governing the
sensitivity of terrestrial ecosystems to nutrient enrichment from N deposition include the
degree ofN limitation, rates of N deposition, elevation, species composition, length of
growing season, and soil N retention capacity. Thus, ecosystems such as alpine tundra,
which are typically strongly N limited, contain vegetation adapted to low N availability,
often have thin soils with limited N retention capacity, and have short growing seasons,
can be particularly sensitive to N deposition. Similarly, the ability of atmospheric N
deposition to override the natural spatial heterogeneity in N availability in arid
ecosystems, such as the Mojave Desert and CSS ecosystems in southern California,
makes these ecosystems more prone to wildfires and sensitive to N deposition.
In the 2008 ISA, effects on individual plant species had not been well studied in the U.S.
More was known about the sensitivity of particular plant communities, based largely on
results obtained in more extensive studies conducted in Europe, which included
hardwood forests, alpine meadows, arid and semiarid lands, and grassland ecosystems.
Among communities sensitive to N deposition, lichens and ectomycorrhizal fungi have
particularly low thresholds for effects (see Appendix 6.5.1 and Appendix 6.5.2). Thus,
the ecosystems containing a large number and/or diversity of these organisms, such as
June 2018
6-161
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
temperate and boreal forests and alpine tundras, could be considered particularly sensitive
to N deposition. More broadly, there has been substantial work on critical loads for N in
U.S. ecoregions since the 2008 ISA (see Appendix 6.5). creating an improved
understanding of which processes, taxa, and regions are sensitive to N deposition
impacts.
6.4. Climate Modification of Ecosystem Nitrogen Response
Biotic responses to N deposition can be modified by climatic shifts in temperature and
precipitation. Appendix 13 provides an overview of this topic, whereas this section
describes in brief some of the climate modifications specific to the N response of
terrestrial ecosystems, particularly plant productivity and diversity. Shifts in temperature
and precipitation can alter the effects of N deposition on these two endpoints, as well as
the N response of microbial communities.
Temperature and precipitation can interact with N to affect plant productivity. As noted
in Appendix 6.2. Xia and Wan (2008) identified almost 1,600 observations of plant
biomass growth in response to N additions, excluding agricultural and horticultural
species (Figure 6-1 A). They found that biomass responses to N increased linearly with
mean annual precipitation (MAP). LeBauer and Treseder (2008) had a considerably
smaller data set (n = 126) than Xia and Wan (2008) and did not find that MAP, mean
annual temperature (MAT), or latitude had significant overall influences on
responsiveness of NPP to added N. However, within individual biomes, forest and tundra
NPP responses to N increased with MAT, and forests also became more responsive with
greater MAP. In general, as precipitation increases, water limitation to plant productivity
is relieved, allowing vegetation to be more responsive to changes in N.
As also described in Appendix 6.2. Tian et al. (2016a) developed a "N response
efficiency" metric (100% * (ANPPtreatment - ANPPControi)/ANPPControi/N addition rate) in
order to examine how plant growth responses per unit N changed with increasing levels
of N. They found N response efficiency significantly changed with environmental
factors. Response efficiency decreased with soil pH and total soil N concentration, and
increased with soil C:N ratios and with precipitation until annual precipitation reached
800 mm/year. The N response efficiency also peaked with moderate annual temperatures
(~8°C) and declined under cooler or warmer conditions (Tian et al.. 2016a).
Not surprisingly, climate modification of plant productivity responses can be particularly
pronounced in drier ecosystems. In the Sonoran Desert near Phoenix, Hall et al. (2011)
added 60 kg N/ha/yr and found little or no increase in production among herbaceous
annuals in low precipitation years, moderate N responses with average rainfall, and
June 2018
6-162
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
strong increases in biomass with added N during above-normal rainfall seasons. Vourlitis
(2012) measured aboveground biomass and litter production in a mature CSS stand in
southern California over an 8-year period. The effect of added N positively correlated
with precipitation and was only significant in the high rainfall years. Similarly ,Zhang_et
al. (2015e) observed greater plant productivity responses to N in years with higher
precipitation in a steppe grassland in Inner Mongolia.
In addition to shifts in plant productivity, plant diversity can also be affected by the
interactions of temperature, precipitation, and N (Porter etal.. 2013). In their
national-scale analysis of herbaceous species richness, Simkin et al. (2016) found that
temperature and precipitation could moderate N effects in some instances. They did not
observe a significant interaction between N and temperature or precipitation on plant
species richness for closed canopy systems (deciduous, evergreen, and mixed forests), but
did in open canopy systems (grasslands, shrublands, and woodlands). In these open
canopy ecosystems, they found that N had a more negative effect on species richness at
lower temperatures. This finding is consistent with Clark et al. (2007) who also found a
small increase in the risk of species loss with lower annual minimum temperatures,
among other factors. Even though the overall effects within closed systems were not
significant, Simkin et al. (2016) also analyzed smaller regional forest gradients
dominated by maple-birch (Acer-Betula), white oak (Quercus alba), or Douglas-fir
(Pseudotsuga menziesii) trees. At this scale, the threshold for negative deposition effects
tended to be lower (7.5-9.5 kg N/ha/yr), and the negative effects of N deposition on
species richness were greatest when temperature and precipitation were high.
Some studies have explored these interactive effects under future climatic and N
scenarios. In ForSAFE-VEG model projections of plant community composition in three
French forests, two N reduction scenarios—the maximum feasible N emission reductions
scenario and the current European legislation scenario for reactive N emissions
rates—resulted in gradual shifts over the next 90 years back toward the plant community
composition observed at the beginning of the 20th century (Rizzetto et al.. 2016).
Notably, the recovery of these plant communities occurred only if climatic factors were
held constant at current levels. In another modeling study, Phelan et al. (2016) made a
similar observation for understory plant community composition in northern hardwood
forests at Bear Brook Watershed in Maine and Hubbard Brook in New Hampshire: the
simulated plant community composition returned toward preindustrial conditions over the
next century only in a scenario in which N deposition rates returned to background and
climate was kept stable.
Finally, in addition to plant responses, climatic factors also can modify the N response of
microbial communities and other soil biota. For instance, N deposition has been shown to
6-163
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
alter the soil microbial community responsible for a step in nitrification, particularly at
ambient temperature and precipitation (Horz etal.. 2004). The N effect was then reduced
when temperature and precipitation were increased (Horz et al.. 2004). Similarly, N
additions and lower precipitation interacted to shift microbial community composition in
forests soils in northeastern China (Wang et al.. 2014c) and in grassland soils in Inner
Mongolia (Li et al.. 2016a). Jarvis et al. (2013) found that both changes in precipitation
and nitrogen were associated with shifts in ectomycorrhizal species composition in
European Scots pine (Pinus sylvestris) stands, although their observational study did not
allow them to test the interactive effects. In a Minnesota grassland, Eisenhauer et al.
(2012) observed a decline in several categories of biota within a soil food web (numbers
of nematode predators, microarthropod herbivores, and taxa richness of nematodes and
microarthropods) under increased N, and a decrease in ciliate protists under high N and
drought. The authors suggested the impact of drought may have been limited by the soil
food web's prior adaption to dry conditions because the study was conducted on a sandy
outwash soil, which dries out quickly in the summer. These studies on disparate biota and
ecosystems suggest climatic factors, namely temperature and precipitation, can interact
with N to affect soil biotic communities. In general, the functional implications to
terrestrial ecosystems of these combined climate-and-N induced shifts in soil biota
remain unclear, although there are notable exceptions [e.g., nitrification; (Horz etal..
2004)1.
Overall, studies investigating the interactive effects of climatic factors and N have been
limited to date (Porter et al.. 2013). Despite this, evidence suggests climatic shifts in
temperature and/or precipitation can alter the effects of N on plant productivity and
diversity, and soil microbial communities and other soil biota.
6.5. Critical Loads
As discussed elsewhere in this ISA (e.g., Appendix 1. Appendix 4. Appendix 5). critical
loads (CLs) are a determination of how much atmospheric deposition can be tolerated by
an environmental system before a significant change occurs. Most commonly, this
approach defines a level of deposition associated with an adverse geochemical or
biological response within an ecoregion. As of the 2008 ISA, most of the CLs developed
for North American ecosystems were for aquatic ecosystems (lakes and streams) in
Canada and the northeastern U.S. Among terrestrial ecosystems in North America, the
2008 ISA identified efforts to develop empirical CLs for western ecosystems in the U.S.,
particularly alpine and arid ecosystems. The reported effect levels ranged from 4 to
5 kg N/ha/yrfor changes in the abundance of individual sensitive alpine plant species, to
20 kg N/ha/yr for community level changes in alpine plant communities. Clark and
June 2018
6-164
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Tilman (2008) calculated the CL for the onset of reduced relative species numbers in
grasslands to be 5.3 kg N/ha/yr with a 95% inverse prediction interval of
1.3-9.8 kg N/ha/yr. A CLof 3.1 kg N/ha/yr was considered protective of lichen
communities in the western U.S. (Fenn et al.. 2008). However, as of 2008, there was no
published CL assessment that spanned the U.S.
A large body of work has been published on CLs for N since the 2008 ISA. Most notably,
the U.S. Department of Agriculture (USDA)-Forest Service (FS) created a detailed
national assessment of empirically developed CLs for the U.S. (Pardo etal.. 2011c).
which has also been summarized into a refereed manuscript (Pardo et al.. 2011a). This
national CL document, Assessment of Nitrogen Deposition Effects and Empirical Critical
Loads (Pardo etal.. 2011c). reports CLs for various biological and biogeochemical
endpoints in 15 terrestrial ecoregions. Additionally, in some cases, more recent CLs have
been published, often falling into the range of CLs identified by the USDA-FS
assessment (Pardo etal.. 2011c) (summarized in Appendix 6.6.3).
As with many published CL estimates, most of the CLs reported in the USDA-FS
assessment were empirically derived, based on ecological changes observed along
atmospheric N deposition gradients or in response to experimental N additions at rates
near ambient deposition. Consequently, the links between N deposition and the measured
response variable are direct for empirically derived CLs, and full process-level
knowledge is not required. There can be a few potential sources of uncertainty with
empirically derived CLs, however. The lack of a full process-based understanding can
make it difficult to extrapolate observed results across space and time. Spatially, variation
in biological and biogeochemical processes imposed by climate, geology, biota, and other
environmental factors may alter the observed deposition-response relationship. This is
particularly problematic in areas where N deposition has received sparse research
attention. Pardo et al. (2011c) reported that for some ecoregions, a single study or very
few studies were available. If the variability of ecosystem response to N deposition across
an ecoregion is not available, the estimated CL for N may be relevant for only a single
ecosystem type or a single subregion within the ecoregion. Whereas atmospheric
deposition responds dynamically to shifts in emissions and weather patterns, ecological
processes react to environmental stress at a variety of timescales. Because ecological
changes can be dependent on a series of underlying processes, there can be a time lag
between deposition and ecological responses. Finally, reference plots at the low end of
the deposition gradient may already have been altered from a "pristine" condition, which
can bias CL estimates upwards. This shift from a background state can be particularly
problematic for highly sensitive indicators, as well as in regions such as the northeastern
U.S. that have long and geographically extensive histories of elevated N deposition. In
6-165
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
the USDA-FS assessment, Pardo et al. (2011c) reported CLs that tended to be higher in
regions experiencing greater rates ofN deposition.
Because environmental factors are large influences on both biogeochemical cycling and
biological processes (Pardo etal.. 2011c). the discussion of terrestrial biological CLs in
this ISA focuses largely on research conducted in North America. Notably, Pardo et al.
(2011c) reported that CLs developed for Europe and China both tended to be higher than
those reported in the U.S., likely because of the higher rates of N deposition in these
regions (Figure 6-4). Further, because a large portion of the variation in CLs is a function
of particular ecological receptors (Pardo et al.. 2011c). the analysis of CL research
presented here compares observations for key receptors (e.g., mycorrhizal fungi, lichens,
herbaceous plants) across ecoregions, focusing on the estimates created by Pardo et al.
(2011c) and other CL research published since 2008.
6-166
DRAFT: Do Not Cite or Quote

-------
Tundra
Lichens
Herbaceous
Taiga
Lichens
Mycorrhiza
Shrublands
Northern
Forests
Lichens
Mycorrhiza
Herbaceous
Coniferous: nitrate leaching
Hardwood: nitrate leaching
Northwestern
Forested
Mountains
Lichens
Mycorrhiza
Herbaceous
Subalpine forest: soil N
Marine West
Coast Forests
Lichens
Eastern
Forests
Lichens
Mycorrhiza
Herbaceous
Forests: foliar N
Forest nitrate leaching
Source: Pardo et al. (2011a).
I—l—l
l—l—l
I I
I I
I I
I United States
I Europe
0	5	10	15	20
Empirical N critical load (kg-ha_1-yr1)
Figure 6-4 Comparison of European and U.S. empirical critical loads for
nitrogen from Pardo et al. (2011a).
6.5.1. Mycorrhiza! Fungi
Mycorrhizal fungi are symbiotic organisms hosted on the roots of many plant species,
with important roles in plant nutrient acquisition, belowground C cycling, and as food
sources for other organisms. As noted in Appendix 62 and Appendix 6.3. mycorrhizal
fungi can be sensitive to added N, responding through changes in physiology and growth
June 2018
6-167
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
(Table 6-2. Table 6-3. and Table 6-8). as well as shifts in species richness and community
composition (Table 6-15 and Table 6-16).
Nitrogen CLs for mycorrhizal fungi range between 5 and 12 kg N/ha/yr in the U.S.
(Figure 6-5; Table 6-24) according to studies assessed in Pardo et al. (2011c) and one
additional study conducted in a southern California CSS ecosystem (Allen et al.. 2016).
Pardo etal. (2011c) documented that N deposition in the range of 5 to 10 kg N/ha/yr can
significantly alter ectomycorrhizal fungi community composition and decrease species
richness in N limited conifer forests. Similarly, N deposition levels of 7.8 to
12 kg N/ha/yr can lead to arbuscular mycorrhizal community changes in CSS ecosystems
in California and grasslands in the Midwest (see Mediterranean California and Great
Plains ecoregions in Figure 6-5). due to declines in spore abundance and root
colonization, and changes in community function. Based on additional analysis, (Pardo et
al.. 2011c) suggested the threshold for N effects on mycorrhizae are even lower because
high background deposition precludes consideration of sites receiving deposition at or
near preindustrial levels. The provisional expert judgment was that CLs for mycorrhizal
diversity for sensitive ecosystem types are 5 to 10 kg N/ha/yr (Table 6-24). Pardo et al.
(2011c) indicated there is high uncertainty in this estimate because few studies had been
conducted at low rates of N deposition. Since the publication of the USDA-FS
assessment, there has been one additional study on CLs for mycorrhizae in the U.S. In
that study, Allen et al. (2016) estimated an N deposition CL of 10-12 kg N/ha/yr for
mycorrhizal biodiversity in southern California CSS ecosystems.
6-168
DRAFT: Do Not Cite or Quote

-------
| Reliable
H Fairly Reliable
Expert Judgement
Empirical Critical Load of
N (kg/ha/yr)
5 Marine West Coast Forests
5 - 7 Northern Forest, Taiga
5-10 Northwest Forested Mountains
2,7 -12 Marine West Coast Forests
| 7.8 - 9.2 Mediterranean California
| 12 Great Plains
CL = critical load; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: the range of CLs reported for mycorrhizal fungi is shown for each ecoregion. The hatch marks indicate increasing level of
uncertainty: no hatch marks for the most certain "reliable" category, single hatching for the "fairly reliable" category, and double
hatching for the "expert judgment" category. The color sequence moves from green toward blue and violet as the CL increases. As
the range of the CL gets broader, the saturation of the color decreases. White areas lack data for CL determination for mycorrhizal
fungi.
Source: Pardo et al. (2011c).
Figure 6-5 Map of critical loads for mycorrhizal fungi by ecoregion in the
U.S.
June 2018
6-169
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Table 6-24 Mycorrhizal critical loads.
Type of
Ecosystem
Critical Load
(kg N/ha/yr)
Biological and Chemical
Effects
Study Species
Reference
California
coastal sage
scrub
10-11
Rapid decline in
mycorrhizal biodiversity
Arbuscular mycorrhizal
fungi
Allen etal. (2016)
Sensitive
ecosystem
types in the
U.S.
5-10
Diversity
Mycorrhizal
Pardo etal. (2011c)
Scots pine
forest in
Scotland
5-10
Community composition
Ectomycorrhizal fungi
Jarvis et al. (2013)
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
6.5.2. Lichens and Bryophytes
Lichens are symbioses between fungi and algae or cyanobacteria and are important
components of plant communities in forests, tundra, and some arid ecosystems. Lichens
contribute to ecosystem function by providing food and habitat for wildlife and affecting
nutrient and hydrologic cycling. Like mycorrhizae, the growth, physiology, and
community composition of lichens is sensitive to atmospheric N deposition (Table 6-5.
Table 6-7. and Table 6-23). The 2008 ISA documented observations thatN deposition
altered lichen community composition at deposition rates of 3 to 8 kg N/ha/yr (Fennet
al.. 2003a). In the San Bernardino Mountains in southern California, up to 50% of lichen
species that occurred in the region in the early 1900s had disappeared by the 1990s (Fenn
et al.. 2003a; Nash and Sigal. 1999). The CL has been calculated for lichen communities
in mixed conifer forests in California at 3.1 kg N/ha/yr (Fenn et al.. 2008). In the Pacific
Northwest, lichen communities have shown evidence of changes in response to increased
N pollution, including decreases in the distribution of sensitive lichen taxa and the
replacement of these taxa with nitrophilous species (Gciscr and Neitlich. 2007).
Since the 2008 ISA, it has become clear that although simple N deposition CL estimates
for changes in lichen community composition can be constructed KMcMurrav et al..
2013; Fenn et al.. 2008). accounting for the effects of climate and other environmental
factors can improve lichen CL estimates (Blctt et al.. 2014; Pardo etal.. 2011c; Geiser et
al.. 2010). providing more refined estimates of CL variability and exceedance. As noted
in Appendix 6.2.3.3. changes in lichen growth are best correlated with integrated
June 2018
6-170
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
measures of total N deposition, rather than with the deposition of a single chemical
species such as HNO3 or NH4+. Lichen responses to N deposition such as increases in
thalli N concentration and shifts in community composition have been most tightly
correlated with colocated canopy throughfall measurements of N-NO3 + N-NH4
(McMurrav et al.. 2013; Root et al.. 2013; Jovan et al.. 2012; Fenn et al.. 2007) and
ambient N concentrations in fine particulates of [NH^SC^ and NH4NO3 (Root et al..
2015; Geiser et al.. 2010).
Lichen community composition and lichen thallus N concentrations shift continuously
with all increments of N addition and the response is generally unimodal. In the latter
case, the fastest rate of change occurs with the earliest increments of N addition and
slows as only tolerant species are left in the community. As N deposition increases,
dominance of lichen communities shift from oligotrophic to eutrophic species (i.e., the
number of oligotrophs decrease and the number of eutrophs increase). At very high
deposition levels, total biodiversity decreases.
Critical loads for lichen range between 0.26 to 10.2 kg N/ha/yr based on Pardo et al.
(2011c) and a number of studies for lichens published since 2008 (Figure 6-6.
Table 6-25). In the USDA-FS CL assessment, Pardo etal. (2011c) documented CLs for
lichens of 1 to 9.2 kg N/ha/yr for Level 1 ecoregions. These CLs were predominantly
based on the shift in community composition toward eutrophic lichen species and away
from oligotrophs. The certainty associated with the lichen CL estimates for each
ecoregion varied considerably, in part because of differences in sampling scheme and
intensity. There were highly reliable CL estimates in the Pacific Northwest and
California, where sampling intensity was high and the linkages between N deposition and
lichen community composition have been well documented. Assessments in the eastern
U.S. are more problematic due to historical and contemporary S emissions and acidifying
deposition. Historical information necessary to identify a "pristine" or "clean" state has
been lacking, making it difficult to determine the CL, and the resulting confidence
associated with the CL was low.
6-171
DRAFT: Do Not Cite or Quote

-------
Empirical Critical Load
of N (kg/ha/yr)
¦ 1 - 3 Tundra, Taiga	\ 1	Uncertainty
11.2 - 3.7 Northwest Forested Mountains, Alaska	~ Reliable
2.5 - 7.1 Northwest Forested Mountains	~ Fairly Reliable
2.7 - 9,2 Marine West Coast Forests	KV1 Expert Judgement
3	North American Deserts
3.1 - 6 Mediterranean California
4-6 Northern Forests
4	- 7 Temperate Sierras
4 - 8 Eastern Temperate Forests
CL = critical loads; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: the range of critical loads reported for lichens is shown for each ecoregion. The hatch marks indicate increasing level of
uncertainty: no hatch marks for the most certain "reliable" category, single hatching for the "fairly reliable" category, and double
hatching for the "expert judgment" category. White areas lack data for CL determination for lichens.
Source: Pardo et al. (2011c).
Figure 6-6
Map of critical loads for lichens by ecoregion in the U.S.
June 2018
6-172
DRAFT: Do Not Cite or Quote

-------
Table 6-25 Lichen critical loads.
Type of
Ecosystem
Critical Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site
Study Species
Reference
Chaparral and
oak woodland
5.5
Shift to nitrophyte dominance
in the lichen community
California
Epiphytic lichens
Pardo etal. (2011c)
Fenn etal. (2010)


Relating total N in throughfall
to lichen biodiversity
Southern California, north of
Los Angeles
Epiphytic lichens
Jovan et al. (2012)
Under etal. (2013)
Forest
0.26 to 0.33 kg inorganic
N/ha/yr and 0.044-0.055
mg/L of ammonium wet
deposition
Declines in presence and
abundance of sensitive lichen
communities
Western Oregon and
Washington
Lichens
Glavich and Geiser (2008)
Forest
<4.1
Poorer thallus condition
Wind River Range, WY,
including the Class I Bridger
Wilderness
Epiphytic lichens
McMurrav et al. (2013)
Forest
4-6 for total N deposition
Decreases in lichen species
richness and N sensitive
species, and poorer thallus
condition
Northeastern U.S. Class I
areas
Epiphytic lichens
Cleavitt et al. (2015)
Forest
4
Degradation to lichen
communities
Northern Rocky Mountains
Epiphytic lichens
McMurrav et al. (2015)
Forest
1.54 and 2.51 kg N/ha/yr
of through-fall dissolved
inorganic N deposition
Lichen communities and lichen
N concentration
Pacific Northwest
Epiphytic lichens
Root etal. (2015)
Mixed conifer
forests
3.1
Enhanced lichen tissue N
concentrations
California
Epiphytic lichens
Fenn et al. (2008)
Fenn etal. (2010)
June 2018
6-173
DRAFT: Do Not Cite or Quote

-------
Table 6 25 (Continued): Lichen critical loads.
Type of
Ecosystem
Critical Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site
Study Species
Reference
Mixed conifer
forests
5.2
Lichen community shifted from
acidophyte dominance to
neutrophyte dominance
California
Epiphytic lichens
Fenn et al. (2008)
Fenn et al. (2010)
Mixed conifer
forests
10.2
Lichen species classified as
acidophytes were extirpated
California
Epiphytic lichens
Fenn et al. (2008)
Fenn et al. (2010)
Temperate forest
3-9
Sensitive species declines of
20-40%
Western Oregon and
Washington
Epiphytic lichens
Geiser et al. (2010)
Various
1-9.2
Lichen health and community
composition
U.S. (national)
Epiphytic lichens
Pardo et al. (2011c)
CL = critical load; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
June 2018
6-174
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
In addition to the USDA-FS assessment by Pardo etal. (2011c). there have been a
number of N deposition CL studies for lichens published since 2008, including research
in the northeastern U.S. (Clcavitt etal.. 2015). California (Fcnn et al.. 2010; Fenn et al..
2008). the Pacific Northwest (Root et al.. 2015; Geiser et al.. 2010; Glavich and Geiser.
2008). and the Rocky Mountains (McMurrav et al.. 2013). In the northeastern U.S.,
Cleavitt et al. (2015) studied four Class I areas and found annual mean and cumulative
deposition of N were strongly negatively correlated with lichen species richness, the
abundance of N sensitive species, and thallus condition. The work by (Cleavitt et al..
2015) supported the CL of 4-6 kg N/ha/yr of total N deposition for epiphytic lichens in
the Northern Forest Ecoregion created by Pardo etal. (2011c).
Based on sampling in mixed conifer forests in the Sierra Nevada and San Bernardino
Mountains in California, Fenn et al. (2008) developed a CL estimate of 3.1 kg N/ha/yr for
throughfall N deposition based on increases in thallus N concentration for the lichen
Letharia vulpina. Above this CL, there were shifts in lichen community composition
away from acidophytic lichen species and toward lichen species considered to be
neutrophytic and nitrophytic. Fenn et al. (2010) later applied these and other previously
published CL values for lichens in California to estimate that 53 and 41% of the chaparral
and oak woodland ecosystems in the state received N deposition in excess of the lichen
CL of 5.5 kg N/ha/yr.
In the Wind River Range in western Wyoming, McMurrav et al. (2013) developed a
lichen CL based on study sites adjacent to the Bridger Wilderness, a Class I area
downwind of an area experiencing oil and gas production. Above a threshold of
4.0 kg N/ha/yr of throughfall N deposition, McMurrav et al. (2013) observed a
degradation of lichen communities in this area that included necrotic and bleached thalli
and decreased growth.
6.5.3. Herbaceous and Shrub Species
Herbaceous species and shrubs are found in grasslands, shrublands, forests, deserts, and
wetlands, and comprise the majority of the roughly 26,600 vascular plant species found
in North America north of Mexico (NRCS. 2009). In forests, herbaceous-layer
(understory) vegetation can be important contributors to ecosystem processes such as
litter production and N cycling (Talhelm et al.. 2013). and can comprise up to 90% of
forest plant biodiversity, including endangered or threatened species (Gilliam. 2007). As
described in Appendix 6.3.3.2. there is abundant evidence that forest understory
vegetation composition can be sensitive to N deposition (Table 6-13).
June 2018
6-175
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Critical loads for herbaceous and shrub species range between 0.9 to 33 kg N/ha/yr based
on studies published since 2008 (Table 6-26). In the USDA-FS assessment, (Pardo et al..
2011c') reported N deposition CLs of 1 to 33 kg N/ha/yr for herbaceous species and
shrubs across all ecoregions (Figure 6-7; Table 6-26). The lowest CL was for tundra
(1-3 kg N/ha/yr), while the highest was for the Mediterranean California ecoregion,
specifically a mixed conifer forest ecosystem in the San Bernardino Mountains (Pardo et
al.. 2011c).
Among the new research published since the USDA-FS assessment was completed,
Simkin et al. (2016) examined the influence of N deposition on species richness of
grasses and forbs in over 15,000 plots located across the continental U.S. Notably, they
found different relationships between N deposition and species richness in open canopy
and closed canopy ecosystems, likely a function of different species loss mechanisms
(Appendix 6.3.2) operating in these systems. In open canopy systems (e.g., grasslands,
shrublands, and woodlands), N deposition above an average of 8.7 kg N/ha/yr led to a
reduction in species richness (5th-95th percentile: 6.4-11.3 kg N/ha/yr). These rates of N
deposition are common across much of the eastern and central U.S., as well as areas in
the western U.S. that are downwind of major urban and agricultural areas (Figure 6-8).
Average CLs for grasses and forbs did not differ widely among the grassland, shrubland,
and woodland ecosystems (8.9, 8.5, and 8.5 kg N/ha/yr). The calculated CL for closed
canopy systems was considerably higher: 13.4 kg N/ha/yr (5th-95th percentile:
6.8-22.2 kg N/ha/yr). At higher levels of N input, processes such as competitive
exclusion and soil acidification occur, decreasing species richness. There are two
important observations from this analysis. First, increases in N deposition below the CL
levels could increase species richness, particularly in open canopy ecosystems. Second,
the effects of N deposition on species richness were often pH dependent (see
Appendix 6.3.2). Simkin et al. (2016) also developed more localized CL estimates. In
forests, these ranged from a low of 7.9 kg N/ha/yr in open canopy eastern forests to a
high of 15.3 kg N/ha/yr in northwestern forest mountains. In the Great Plains region,
(Simkin et al.. 2016) estimated CLs for grasses and forbs of 8.3-9.8 kg N/ha/yr in the
open canopy systems to 11.3-19.6 kg N/ha/yr in closed canopy systems. In arid
ecosystems, Simkin et al. (2016) estimated a CL of 8.3-9.9 kg N/ha/yr for open canopy
ecosystems and 13.5-17 kg N/ha/yr for closed canopy ecosystems.
6-176
DRAFT: Do Not Cite or Quote

-------
Table 6-26 Herbaceous and shrub critical loads.
Type of
Ecosystem
Critical Load
(kg N/ha/yr)
Biological and Chemical
Effects
Study Site
Study Species
Reference
Alpine tundra
3.0
Protection of natural community
cover
Rocky Mountain
National Park
Alpine grasses and forbs
Bowman et al. (2012)
Alpine and	1 to 2	ForSAFE-VEG	Northern and central Alpine and subalpine ground Sverdrup et al. (2012)
subalpine	modeled changes in alpine and Rocky Mountains vegetation, including two tree
ground	subalpine plant community	species (Engelmann spruce and
vegetation	(modeled from 1750 to 2500)	White spruce-P/cea engelmannii
and Picea glauca)
Desert
2.1 and 3.6
Exponential increase in the
Southern CA Joshua
Creosote bush (Larrea tridentata)
Rao et al. (2010)


probability of biomass
Tree National Park
other shrubs, forbs, and grasses;



(simulated using the DayCent

two tree species included



model) exceeding the fire

(California juniper and Single leaf



threshold of 1,000 kg/ha

pinyon-Juniperus californica and





Pinus monophylla)

North
Open: 8.3-9.9
Decreasing species richness
Ecoregion
Various
Simkin et al. (2016)
American
(mean = 9.2,
grasses and forbs



desert
n = 240)





Closed: 13.5-17.0





(mean = 16.5,





n = 32)




Semiarid
<11
Conversion to exotic grasslands
Riverside County, CA
Various
Cox et al. (2014)
coastal sage





scrub





Forest
16.8
Plant biodiversity loss as
The Netherlands
EUNIS classes
De Vries et al. (2010)


predicted by SMART2



Beech and	5-11	Plant biodiversity loss predicted Aeshau, Switzerland Forest understory plants	Belvazid et al. (2011b)
fir/spruce forest	by ForSAFE-VEG
June 2018
6-177
DRAFT: Do Not Cite or Quote

-------
Table 6 26 (Continued): Herbaceous and shrub critical loads.
Type of
Ecosystem
Critical Load
(kg N/ha/yr)
Biological and Chemical
Effects
Study Site
Study Species
Reference
Boreal forests
0.9-7.8
Plant biodiversity loss predicted
by ForSAFE-VEG
Sweden
Forest understory plants
De Vries et al. (2010)
Subalpine
vegetation
1.9-3.5
ForSAFE-VEG modeled
changes in subalpine plant
community between 2010 and
2100
Rocky Mountain
National Park
Subalpine plant community,
consisting primarily of forest
understory plants and one
additional tree species (Subalpine
fir-Abies lasiocarpa)
McDonnell etal. (2014a)
Pine forest
4-6
Plant biodiversity loss predicted
by ForSAFE-VEG
Sostared, Sweden
Forest understory plants
Belvazid et al. (2011b)
Spruce forest
10-16
Plant biodiversity loss predicted
by ForSAFE-VEG
Bachtel, Switzerland
Forest understory plants
Belvazid et al. (2011b)
Spruce forest
1-2
Plant biodiversity loss predicted
by ForSAFE-VEG
Hogbranna, Sweden
Forest understory plants
Belvazid et al. (2011b)
Eastern
temperate
forests
Open: 6.6-9.7
(mean = 7.9,
n = 947)
Closed: 7.8-19.3
(mean = 12.5,
n = 7,378)
Decreasing species richness
grasses and forbs
Ecoregion
Various
Simkin et al. (2016)
Marine West
Coast forests
Open: no data
Closed: 10.4-15.0
(mean = 12.8,
n =24)
Decreasing species richness
grasses and forbs
Ecoregion
Various
Simkin et al. (2016)
Northern
Open: 8.0-9.8
Decreasing species richness
Ecoregion
Various
Simkin et al. (2016)
forests
(mean = 8.9,
n = 75)
Closed: 8.0-18.9
(mean = 13.8,
n = 1,955)
grasses and forbs



June 2018
6-178
DRAFT: Do Not Cite or Quote

-------
Table 6 26 (Continued): Herbaceous and shrub critical loads.
Type of
Ecosystem
Critical Load Biological and Chemical
(kg N/ha/yr) Effects
Study Site
Study Species
Reference
Northwestern
forested
mountains
Open: 8.0-10.2 Decreasing species richness
(mean = 9.1, grasses and forbs
n = 1,429)
Closed: 10.8-19.6
(mean = 15.3,
n = 2,113)
Ecoregion
Various
Simkin et al. (2016)
Temperate	Open: 8.6-8.7 Decreasing species richness Ecoregion	Various	Simkin et al. (2016)
sierras	(mean = 8.65, grasses and forbs
n = 3)
Closed: 14.8-14.8
(mean = 14.8,
n =42)
Dry and neutral
grasslands
8.0
Plant biodiversity loss as
predicted by SMART2
The Netherlands
Drv and neutral qrasslands De Vries et al. (2010)
Semidry
calcareous
grasslands
12.4
Plant biodiversity loss as
predicted by SMART2
The Netherlands
Semidrv calcareous arasslands De Vries et al. (2010)
Moist and wet
oligotrophic
grasslands
12.6
Plant biodiversity loss as
predicted by SMART2
The Netherlands
Moist and wet oliqotrophic De Vries et al. (2010)
grasslands
Great Plains
Open: 8.3-9.8
(mean = 9.3,
n =618)
Closed: 11.3-19.6
(mean = 16.6,
n = 274)
Decreasing species richness
grasses and forbs
Ecoregion
Various Simkin et al. (2016)
Various
1-33
Change in plant community
composition; other various
effects
U.S. (national)
Various Pardo et al. (2011c)
EUNIS = European Nature Information System; ForSAFE-VEG = a dynamic forest ecosystem model; ha
and nutrient-cycling model; yr = year.
= hectare; kg = kilogram; N = nitrogen; SMART2 = a simple soil acidification
1
June 2018
6-179
DRAFT: Do Not Cite or Quote

-------
S-2E Great Plains.	f\^j Falrty Reiiaoie
6 isiga	^ Expert Judgment
6 ¦ 33 Mediterranean California
>7 - <21 Northern Forests
«17.5 Eastern Temperate Fiyeste
CL = critical load; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: the range of CLs reported for herbaceous plants and shrubs is shown for each ecoregion. The hatch marks indicate
increasing level of uncertainty: no hatch marks for the most certain "reliable" category, single hatching for the "fairly reliable"
category, and double hatching for the "expert judgment" category. White areas lack data for CL determination for herbaceous
species and shrubs.
Source: Pardo et al. (2011c).
Figure 6-7 Map of critical loads for herbaceous plants and shrubs by
ecoregion in the U.S.
June 2018
6-180
DRAFT: Do Not Cite or Quote

-------
N Deposition	Critical Load
(kg/ha/yr)	(kg/ha/yr)
H 0-5	CtoMd Open
~	5-9	• 7.4 - 9 ^
~	9-11	• 9-11 A
¦	11-13	11-13
¦	13-15	• 13-15
¦	15-63	• 15-19.6
0	250 500 1.000 Kilometers
1	i i i I i i i I
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: the 3,317 open sites (combined grassland, shrubland, and woodland vegetation types) are portrayed with triangles, and the
11,819 closed canopy sites (deciduous, evergreen, and mixed forests) are portrayed with circles. Background deposition values are
the average of 27 years of wet deposition (NADP 1985-2011) plus the average of 10 years of dry deposition (from Community
Multiscale Air Quality model, 2002-2011). Other variation in CLs is due to the other predictor variables (soil pH, temperature, and
precipitation).
Source: Simkin et al. (2016).
Figure 6-8 Nitrogen deposition (gray scale) and critical loads for nitrogen
deposition based on total graminoid plus forb species richness
(colored symbols).
1
2	Within the Colorado Rocky Mountains, there were several efforts to develop CLs based
3	on both empirical and modeled data. In the Rocky Mountain National Park, Bowman et
4	al. (2012) calculated an empirical CL of 3 kg N/ha/yr to protect natural community cover
5	based on large increases in the abundance of the sedge Carex nipestris to additional N
6	deposition. McDonnell et al. (2014a) applied the ForSAFE-VEG model to develop a
7	long-term CL estimate aimed at avoiding future (2010-2100) changes (of more than
8	10%) in subalpine plant biodiversity in Rocky Mountain National Park. The estimated
June 2018
6-181
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
CL to protect future plant diversity was 1.9-3.5 kg N/ha/yr, a value already exceeded in
the study area. Sverdrup et al. (2012) also used the ForSAFE-VEG model to understand
how long-term CLs would be influenced by climate change, but worked from a synthetic
alpine and subalpine vegetation data set developed from observations at national parks in
the northern and central Rocky Mountains region of the U.S. They determined CL values
of 1 to 2 kg N/ha/yr to protect against a future change in plant diversity of 5-20%.
Notably, CLs protecting against future biodiversity changes may not be comparable
directly to other CLs protective against current changes to herbaceous and shrub
biodiversity (Table 6-26). Critical loads are expressed on an annual basis, and CLs
protective of future biodiversity may be lower generally than shorter-term CLs to avoid
the accumulating effects of deposition over longer periods of time. Consideration of
future temperature increases, among other factors, however, may complicate this
relationship. In their simulations, (McDonnell et al.. 2014a') concluded that where the CL
was between 1.9 and 3.5 kg N/ha/yr depended upon future temperatures. Higher
simulated temperatures resulted in CL values higher in the range, closer to
3.5 kg N/ha/yr. They attributed this to the increased N uptake of the vegetation with
higher temperatures, allowing an increase in deposition while protecting against shifts in
plant diversity. Thus, among other factors, comparison of CLs protecting against future
versus current changes likely requires a consideration of the underlying assumptions
about future conditions.
In addition to the CL estimates for plant species richness (Simkin et al.. 2016). two other
CL estimates have been developed for arid and semiarid ecosystems. Pardo etal. (2011c)
determined a CL of 3-8.4 kg N/ha/yr for North American deserts based on a decrease in
native forbs. Rao et al. (2010) developed unique CL estimates for the rate ofN deposition
that would increase the probability of vegetation productivity in two arid ecosystems
within Joshua Tree National Park in southern California exceeding a wildfire risk
threshold. To do this, Rao et al. (2010) applied the DayCent model. The risk of exceeding
the wildfire risk threshold of 1,000 kg of biomass/ha increased rapidly above CLs of 2.1
and 3.6 kg N/ha/yr for the two ecosystem types. The risk of exceeding the threshold
increased until N deposition levels reached 5.5 and 8.8 kg N/ha/yr for the two ecosystem
types. Notably, contemporary rates ofN deposition at the study sites were 3 to
8 kg N/ha/yr, and up to 16 kg N/ha/yr has been observed in areas adjacent to these desert
ecosystems and downwind from or near urban areas (Fenn et al.. 2010). In southern
California, Cox et al. (2014) modeled the influence of environmental factors such as
climate, land use, and N deposition on the conversion of CSS ecosystems to exotic annual
grasslands from 1930 to 2009. Here, conversion of CSS to exotic grasslands was more
likely to occur if N deposition exceeded 11 kg N/ha/yr. This CL estimate was similar to a
CL of 10 kg N/ha/yr for CSS developed by Pardo et al. (2011c).
6-182
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
6.5.4. Trees
As noted in Appendix 6.3.3.1. there are relatively few direct observations of overstory
tree community composition change, likely because it is difficult to observe changes for
long-lived organisms in slowly developing communities. However, there is more
abundant evidence of changes in tree growth, mortality, and physiology
(Appendix 6.2.3.1'). Notably, CLs for trees due to N + S deposition are presented in
Appendix 5.
Nitrogen CLs for trees range between >3 to 39 kg N/ha/yr based on studies compiled in
Pardo etal. (2011c) and one other study published since 2008 (Table 6-27). In the
USDA-FS assessment, Pardo etal. (2011c) reported that N CLs for forest ecosystems
ranged from >3 to 39 kg N/ha/yr (Figure 6-9). Estimates of 3-26 kg N/ha/yr for declining
forest growth and increased mortality were based on the forest inventory data analysis
conducted by Thomas et al. (2010) and a chronic N addition experiment conducted in red
spruce forests in the northeastern U.S. (McNulty, 2005, 092173). In California mixed
conifers, Pardo etal. (2011c) reported a CL of 17 kg N/ha/yr for decreased fine root
biomass and 39 kg N/ha/yr for forest sustainability, both developed from observations
reported by Fenn et al. (2008). In one additional study subsequent to the USDA-FS
assessment, Fleischer et al. (2013) reported that greater rates of N deposition increased
canopy photosynthetic capacity in evergreen needleleaf forests below a CL threshold of
8 kg N/ha/yr.
June 2018
6-183
DRAFT: Do Not Cite or Quote

-------
Table 6-27 Tree critical loads.
Type of Critical Load
Ecosystem kg N/ha/yr
Biological and Chemical
Effects
Study Site
Study
Species
Reference
Evergreen ~8
Saturation of
32 forest sites
Conifer
Fleischer et al.
forest
photosynthetic capacity of
around the globe

(2013)

the canopy



Various >3-39
forests
Tree growth and survival;
other various effects
U.S. (national)
Various
Pardo et al. (2011c)
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
June 2018
6-184
DRAFT: Do Not Cite or Quote

-------
Empirical CL
>3-S
>3-<26
4-17
<5-10
¦ 5
17-30
of N (kg ha"' yr
Easter^ Temperate
Northern Forests
Northwest Forested Mountains
Tropical and Subtropical Humid Forests
Marine West Coast Forests
Mediterranean Catttamia
|\J Fairly Reliah'e
Expert Judgnen
CL = critical load; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: the range of CLs reported for forest ecosystems is shown for each ecoregion; this map does not include the responses of
mycorrhizal fungi, lichens, or understory herbaceous plants already represented. The hatch marks indicate increasing level of
uncertainty: no hatch marks for the most certain "reliable" category, single hatching for the "fairly reliable" category, and double
hatching for the "expert judgment" category. White areas lack data for CL determination for forest ecosystems.
Source: Pardo et al. (2011cl
Figure 6-9 Map of critical loads for forest ecosystems by ecoregion in the
U.S.
June 2018
6-185
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
6.5.5.
Critical Loads Exceedance Studies
In the USDA-FS assessment ofN deposition CLs, Pardo et al. (2011c) evaluated the CL
exceedances for mycorrhizal fungi, lichens and bryophytes, herbaceous plants and
shrubs, and forests using N deposition estimates produced by the Community Multiscale
Air Quality (CMAQ) model v.4.3. This CMAQ model used data reported in 2001 for the
simulations of wet plus dry N deposition. Pardo etal. (2011c) concluded that large parts
of the eastern U.S., as well as localized areas in the West, experience rates of N
deposition that exceed the CL for sensitive ecosystem components. In this assessment,
Pardo etal. (2011c) determined that the resources most threatened by elevated N
deposition included freshwater diatoms, lichens, bryophytes, and herbaceous plants.
In addition to Pardo etal. (2011c). several studies have quantified N deposition CL
exceedance at the regional scale for herbaceous plants. In the Simkin et al. (2016)
analysis of herbaceous species richness in over 15,000 plots over the continental U.S.,
approximately 41% of grassland plots were experiencing N deposition in exceedance of
the CL for plant species richness. Clark et al. (2013) used 26 years (1985-2010) of N
deposition data with ecosystem-specific functional responses from local field
experiments and a national CLs database to generate estimates of herbaceous species
loss. In scenarios using the low end of the CL range, N deposition exceeded CLs over
0.38, 6.5, 13.1, 88.6, and 222.1 million ha for the Mediterranean California, North
American Desert, Northwestern Forested Mountains, Great Plains, and Eastern Forest
ecoregions, respectively, with corresponding species losses ranging from <1 to 30%.
When scenarios assumed less sensitivity (using a common CL of 10 kg N/ha/yr, and the
high end of the CL range) minimal losses were estimated. The large range in projected
impacts among scenarios highlights the uncertainty in current CLs estimates for plant
diversity in the U.S.
McLauchlan et al. (2014) analyzed a 27-year record of ecophysiological, community, and
ecosystem metrics for an annually burned Kansas tallgrass prairie. Despite observed rates
of atmospheric N deposition (7 kg N/ha/yr from 2002 to 2009) that exceed the minimum
estimated CL for herbaceous plants in tallgrass prairie of the Great Plains of North
America [5-15 kg N/ha/yr; (Pardo et al.. 201 lc)l. there were no observations of plant
community composition change or effects on plant physiology that signaled an obvious
influence ofN deposition: plant N concentrations and plant N availability did not
increase, aboveground NPP was unchanged, forb diversity did not decline, and the
relative abundance of dominant grasses did not shift toward more eutrophic species.
Thus, current rates ofN deposition do not appear to be altering ecosystem function in this
grassland, even though these rates exceed minimum CLs. The authors offered a couple of
potential explanations for this lack of response, including sequestration of the additional
June 2018
6-186
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
N into soil organic matter pools and offsetting removal of N by biomass burning and/or
grazing.
Finally, Clark et al. (2018) estimated CL exceedance areas for the conterminous U.S.
over a more than 200 year period. In this study, they examined six CL types, with three N
CLs germane to terrestrial ecosystems: (1) changes in forest tree health, (2) changes in
lichen communities, and (3) changes in herbaceous and shrub plant community
composition. Overall, this analysis showed terrestrial N CLs have been exceeded for
many decades in areas across the U.S. Minimum values for these CLs were already
exceeded by 1855 for the first two CLs, and the latter CL was exceeded between 1935
and 1955. Exceedance areas peaked in 1995 for changes in lichen communities and plant
community composition at 3.47 and 2.87 million km2, respectively, before declining
marginally by 2006. The minimum forest tree health CL was exceeded in 2.41 million
km2 by 1855 and did not change much over time, primarily because the relatively low CL
compared to deposition values in the Eastern Temperate and Northern forest ecoregions.
6.6. Summary
6.6.1. Physiology, Growth, and Productivity Summary
At the time of the 2008 ISA, there was evidence sufficient to infer a causal relationship
between N deposition and the alteration of the biogeochemical cycling of C in terrestrial
ecosystems. This evidence included observations that added N increased plant
productivity across abroad range of terrestrial ecosystems. The 2008 ISA, however,
made no explicit statement regarding the effects of N deposition on physiology, growth,
and productivity.
Since the 2008 ISA, a more complete understanding of how N deposition effects
terrestrial organisms and ecosystems has been developed, the effects of N deposition in
terrestrial ecosystems in North America and Europe have been more widely documented,
and a significant new body of research has observed N deposition impacts in Asia
(particularly China). These efforts have provided further evidence that added N alters
plant physiological processes, stimulates the growth of most plants, and broadly increases
productivity (Liu and Greaver. 2010; LeBauer and Treseder. 2008; Xia and Wan. 2008).
Moreover, studies have shown species-specific effects of N deposition on tree growth and
mortality (Dietze and Moorcroft. 2011; Thomas et al.. 2010). There is also widespread
evidence that N additions affect soil microbial physiology and biomass (see Appendix 4
for soil extracellular enzyme responses), particularly by decreasing the abundance of
June 2018
6-187
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
symbiotic mycorrhizal fungi (Li et al.. 2015; Treseder. 2008). There have also been
numerous observations that additional N can alter arthropod performance and abundance
among both detritivores and herbivores I (Throop and Lerdau. 2004); Table 6-171. Thus, it
is apparent that plants, microorganisms, and ecosystem-scale productivity are sensitive to
N availability. This body of evidence is sufficient to infer a causal relationship
between N deposition and the alteration of the physiology and growth of terrestrial
organisms and the productivity of terrestrial ecosystems.
Although plants are broadly sensitive to added N, it is clear that the effects of N on plant
growth and productivity vary considerably among plant tissues (Li et al.. 2015; Xia and
Wan. 2008). individual species and functional types (Xia and Wan. 2008). and biomes
(LeBauer and Treseder. 2008). In a meta-analysis of 1,600 observations, short-lived plant
functional types tended to have stronger relative responses than long-lived plants;
herbaceous plants responded more than woody plants, and annual herbs responded more
than perennial herbs IYXia and Wan. 2008); Figure 6-11. Shrubs were notably less
responsive than trees (Figure 6-1). potentially because of the prevalence of shrubs in
more arid environments; overall, biomass responses increased linearly with mean annual
precipitation (Xia and Wan. 2008). There is also now more consistent and widespread
evidence that N additions increase aboveground biomass more than belowground
biomass, raising the shoot:root ratio among plants (Li et al.. 2015; Lu et al.. 2011b).
The increase in aboveground growth of individual plant species is the best documented
response of plants to N, with fewer studies documenting belowground growth responses
and more complex physiological and ecosystem-scale responses (e.g., Figure 6-1 and
Figure 6-2). LeBauer and Treseder (2008) found that N additions stimulated aboveground
productivity, but the relative stimulation was smaller than the response of individual
plants identified by Xia and Wan (2008). Belowground net primary productivity
responses to added N have not yet been synthesized because of lack of data (LeBauer and
Treseder. 2008). Liu and Greaver (2009) did not find a significant effect on net
ecosystem exchange, but forest ecosystem C content was increased by 6%.
In the 2008 ISA, the increase in plant growth and terrestrial productivity was attributed to
increases in foliar N content and related increases in photosynthesis. As noted earlier,
increases in plant N concentrations are widespread (Koricheva et al.. 1998). particularly
in the foliage. Koricheva et al. (1998) had previously quantified increases in foliar N
concentrations among woody plants; Xia and Wan (2008) expanded this analysis to other
plant functional groups, with an overall stimulation of +28.5%. Aside from smaller
responses among legumes and broader functional groups containing legumes, there was a
relatively narrow range of responses across functional groups (+24-34%). Root N
concentrations show similar responses (Li et al.. 2015; Xia and Wan. 2008). Given the
6-188
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
mechanistic and empirical relationships between greater foliar N concentrations, higher
leaf concentrations of the C assimilating enzyme RubisCo (Evans. 1989). and greater
maximum rates of leaf-level photosynthesis in vascular plants (Wright et al.. 2004). the
increases in foliar N observed with greater N supply have been linked to higher rates of
leaf-level photosynthesis [e.g., (Teskev et al.. 1994)1. Alternately, there was evidence in
the 2008 ISA that much of the additional foliar N could be physiologically inactive
(Bauer etal.. 2004). consistent with the significant increase in free amino acids observed
by Koricheva et al. (1998). Increases in photosynthesis following N additions have been
observed across a variety of plant functional types, but higher rates of photosynthesis
have not been universally observed in long-term experiments [e.g., (Talhelm et al.. 2011;
Elvir et al.. 2006; Chen et al.. 2005b; Newman et al.. 2003; Laitha and Whitford. 1989;
Gulmon and Chu. 1981)1 and there have not been any broad syntheses or meta-analyses
of either leaf-level photosynthesis or gross primary productivity. There is a similar
mechanistic and empirical relationship between tissue N concentrations and respiration
rates, but this relationship has shifted in long-term N addition studies, providing further
evidence that much of the additional tissue N is physiologically inactive [e.g., (Burton et
al.. 2012; Drake et al.. 2008)1.
Decreases in the quantity of C allocated belowground to roots, mycorrhizae, and root
exudation could provide a mechanism that increases aboveground plant growth in plants
and ecosystems that do not show gains in photosynthesis [e.g., (Talhelm et al.. 2011)1.
Plants provide mycorrhizal fungal symbionts with C in exchange for nutrients, and the
abundance of mycorrhizal fungi and the rates at which these fungi colonize roots declines
when N is added to terrestrial ecosystems (Treseder. 2004). a finding that has since been
corroborated (Li et al.. 2015; Treseder. 2008). The decrease is particularly well
documented for ectomycorrhizal fungi (Table 6-2). but can also occur with arbuscular
mycorrhizal fungi IYvan Diepen et al.. 2010); Table 6-31.
Nitrogen deposition can affect the physiology and abundance of soil microorganisms
through decreases in soil pH, increases in inorganic N availability, changes in soil food
webs, and shifts in the quantity and quality of available C (Xia etal.. 2015; Manning et
al.. 2008; Treseder. 2008; Koricheva et al.. 1998). There were some observations in the
2008 ISA that added N decreased microbial biomass. There is now more abundant
evidence that N deposition can affect microbial communities. In meta-analyses, N
additions have been shown to decrease microbial biomass (Treseder. 2008). microbial
biomass C (Liu and Greaver. 2010). and microbial biomass N (Luetal.. 2011b). In a
meta-analysis, Treseder (2008) found that the negative effects of added N on microbial
biomass increased in magnitude with the duration of the N additions and the cumulative
amount of added N. Notably, these observations of soil microbial biomass would include
the hyphal biomass of mycorrhizal fungi. There are fewer observations of N addition
6-189
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
effects at the level of individual microbial domains (e.g., bacteria, fungi), functional
groups (e.g., saprotrophic vs. mycorrhizal fungi), or other higher-level taxonomic groups,
and aside from the aforementioned decreases in mycorrhizal fungi, effects at these scales
appear to be less consistent (Table 6-4).
6.6.2. Biodiversity Summary
The 2008 ISA found evidence sufficient to infer a causal relationship between deposition
and the alteration of species richness, species composition, and biodiversity in terrestrial
ecosystems. Since 2008, there is now more widespread documentation of decreases in
lichen species richness as the result of N deposition in the U.S. [e.g., (Geiser et al.. 2010;
Jovan. 2008)1. and there are now direct observations that: (1) N deposition in the U.S. is
altering herbaceous plant species richness across a broad range of ecosystems, including
forests, grasslands, arid and semiarid ecosystems, and alpine tundra [e.g., (Simkin et al..
2016)1. and (2) N deposition in the U.S. is changing mycorrhizal community composition
[e.g., (Allen et al.. 2016; Lilleskov et al.. 2008)1. Further, based on changes in mortality
and growth rates of overstory tree species, there is also now indirect evidence that N
deposition is altering overstory tree community composition [e.g., (Dietze and Moorcroft.
2011; Thomas et al.. 2010)1. The body of evidence is sufficient to infer a causal
relationship between N deposition and the alteration of species richness, community
composition, and biodiversity in terrestrial ecosystems.
In forests, research since 2008 has provided evidence of altered understory plant, soil
microbial, arbuscular and ectomycorrhizal, and arthropod communities. Two new
assessments using forest inventory plots from the eastern U.S. have observed effects on
growth and mortality that likely impact forest overstory tree composition [e.g., (Dietze
and Moorcroft. 2011; Thomas et al.. 2010)1. However, these data sets and analyses have
not directly been applied to assess changes in forest overstory composition either in
North America or Europe. Evidence for altered forest understory plant communities (also
known as herbaceous layer or groundcover vegetation) come from both the 2008 ISA and
the literature published since 2008. Gilliam (2006) reviewed nine studies on the effects of
N deposition on forest understory plant communities in North America and Europe,
including two European studies that documented shifts in understory plant community
composition along N deposition gradients (Strengbom et al.. 2001; Brunetetal.. 1998). In
a new analysis of understory plant community composition, Simkin et al. (2016)
observed that N deposition increased herbaceous plant species richness where soil pH
was neutral or basic, but that N deposition above 11.6 kg N/ha/yr decreased herbaceous
species richness on acidic soils. Similar studies of plant community composition have
been conducted across N deposition gradients in Europe, with both Verheven et al.
June 2018
6-190
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
(2012) and Dirnbock et al. (2014) finding shifts in plant community composition toward
more nutrient-demanding and shade-tolerant plant species, but no loss of species
richness. For soil microbial communities, Zechmeister-Boltenstern et al. (2011) observed
shifts in soil microbial community composition along an N deposition gradient in Europe
and multiple N addition studies identified since 2008 have observed changes in microbial
community composition (e.g., (Zhao et al.. 2014a; Hobbie et al.. 2012; van Diepen et al..
2010). Among studies of how N affects forest ectomycorrhizal community composition,
there were changes in six out of seven studies identified (Table 6-15). This included four
studies where shifts in community composition could be directly or indirectly linked to N
deposition gradients in the U.S. or Europe.
Since 2008, new studies have quantified the impact of N additions on species richness,
species diversity, and community composition among vascular plants, bryophytes,
lichens, and soil microorganisms in alpine and Arctic tundra ecosystems in North
America, Europe, and Asia. Within the U.S., several N addition studies have documented
changes in plant community composition, including experiments in Colorado (Farrer et
al.. 2015; Bowman et al.. 2012) and Washington (Bishop et al.. 2010). In the North
Atlantic region of Europe, Armitage et al. (2014) observed shifts in community
composition and decreases in plant species richness among alpine heathlands located
along an N deposition gradient. Similarly, Southon et al. (2013) found decreasing
vascular plant species richness with increasing N deposition in heathlands in the U.K.
Arens et al. (2008) did not observe changes in community composition as a result of N
additions in Greenland, but shifts in plant community composition were observed in
longer (7 to 8 years) experiments in tundra ecosystems in China (Song and Yu. 2015).
Switzerland (Bassin et al.. 2013). and Sweden (Wardle et al.. 2013). Shifts in microbial
community composition were observed in several N addition experiments, including in
the U.S. in Colorado (Farrer etal.. 2013; Nemergut et al.. 2008).
New research on lichens has added further evidence that lichen communities in the U.S.
and Europe are sensitive to current levels of atmospheric N deposition. Shifts in lichen
community composition attributable to atmospheric N pollution have been observed in
forests throughout the West Coast (Jovan et al.. 2012; Geiser et al.. 2010; Jovan. 2008).
in the Rocky Mountains (McMurrav et al.. 2015; Rogers et al.. 2009). and in southeastern
Alaska (Schirokauer et al.. 2014b). Outside of the U.S., changes in lichen community
composition have been observed with increased atmospheric NO2 concentrations in Nova
Scotia, Canada (Gibson et al.. 2013) and experimental N additions in Sweden (Johansson
etal.. 2012).
Since 2008, there have been direct observations of reduced species richness along
atmospheric N deposition gradients for grasslands in the U.S. (Simkin et al.. 2016) and in
6-191
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Europe (Pannck et al.. 2015; Field et al.. 2014; Gaudnik et al.. 2011; Stevens et al..
2011a; Stevens et al.. 201 lb; Dupre et al.. 2010; Stevens et al.. 2010b; Stevens et al..
2010a'). Studies in both the U.S. and in Europe have found that the effects of N deposition
on grassland communities are often dependent on soil pH [e.g., (Simkin et al.. 2016;
Stevens et al.. 201 la; Maskell et al.. 2010; Stevens et al.. 2010b)l. Information about
changes in mycorrhizal communities in grasslands was limited and provided mixed
results (Chen et al.. 2014); while of the five studies identified investigating compositional
changes to soil microbial communities in grasslands, four observed shifts in composition
(Table 6-20).
Research since 2008 from N deposition gradient studies and N addition experiments in
U.S. arid and semiarid regions has been dominated by research from southern California,
which has provided evidence that higher N availability alters plant community
composition through an increase in invasive annual plants (Cox et al.. 2014; Allen et al..
2009; Rao et al.. 2009). Many of these studies documented changes in plant community
composition, with fewer observations of plant species loss. Evidence for shifts in plant
community composition is particularly strong for CSS and chaparral ecosystems near the
southern California coast and in areas of the Mojave Desert downwind of major southern
California population centers. An important effect of N deposition in arid ecosystems is
an increase in N availability in the interspaces between shrubs, allowing the growth of
annual grasses and forbs, which can grow and reproduce during the brief seasonal periods
with adequate moisture availability (Brooks. 2003). This phenomenon provides a more
continuous fuel bed for wildfires, increasing fire frequency and shifting plant community
composition away from species that are not fire-adapted (Padgett and Allen. 1999; Allen
et al.. 1998). Relative to plants, there are fewer studies of N effects on microbial
biodiversity in arid and semiarid ecosystems, with these studies showing mixed results
(Table 6-22).
6.6.3. Critical Loads Summary
The 2008 ISA documented efforts to develop CLs in the U.S. However, the CLs for
terrestrial ecosystems available in 2008 were for a subset of western ecosystems. There
were no published assessments of N deposition CLs spanning ecosystems across the U.S.
Since the 2008 ISA, there has been substantial work on CLs for N for U.S. ecoregions. A
large body of work has been published on CLs for N since the 2008 ISA. Notably, the
USDA-FS published Assessment of Nitrogen Deposition Effects and Empirical Critical
Loads (Pardo etal.. 2011c). which reports CLs for various biological and biogeochemical
endpoints in terrestrial ecoregions (Omernick Level 1) in the U.S.
June 2018
6-192
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Most of the published CLs included in the Pardo etal. (2011c) assessment or published
subsequently have covered lichens, mycorrhizae, and herbaceous and understory plant
species. There are still relatively few observations of CLs for overstory tree species.
There do not appear to have been any CLs published for birds, mammals, or arthropods
in terrestrial ecosystems. In general, CLs were higher in regions that experienced greater
rates of ambient N deposition. In part, these higher critical loads may represent previous
ecological change caused by historical N deposition. This pattern would explain why the
empirical CL is often above ambient deposition even as that deposition increases in the
same ecosystem type across a region (Pardo et al.. 2011c). The CLs published since the
publication of the USDA-FS assessment by Pardo etal. (2011c) often fall into the range
of CLs identified by (Pardo et al.. 2011c). particularly if these CLs assess similar
ecological endpoints. The new information is presented in tandem with the CLs by Pardo
et al. (2011c) in Table 6-28. Notably, the CLs reported for Pardo et al. (2011c) are a
range for the entire Level 1 ecoregion, whereas the new CL estimates are either for the
entire ecoregion or specific ecosystems within each respective ecoregion.
6-193
DRAFT: Do Not Cite or Quote

-------
Table 6-28 Critical loads for nitrogen by Pardo et al. (2011c) and more recent
critical load information.
The critical loads (CLs) reported for Pardo et al. (2011c) are for Level 1
Ecoregions, whereas the new CL information may be for more specific
ecosystems within each respective Ecoregion (e.g., California coastal sage scrub
within the Mediterranean California Ecoregion; this is noted in the 5th column of
the table entitled "Type of Ecosystem").
Pardo etal. (2011c)
New Critical Load Information
Upper
Lower Critical
Critical Load
Ecoregion Load (kg (kg
(Level 1) N/ha/yr) N/ha/yr)
Biological
and
Critical Load Type of Chemical Study
(kg N/ha/yr) Ecosystem Effects Species Reference
Mycorrhizal fungi
Mediterranean 7.8 9.2
California
10-11 California Rapid decline Arbuscular Allen et al. (2016)
coastal sage in mycorrhizal mycorrhizal
scrub biodiversity fungi
North n/a n/a
American
deserts
n/a
Southern n/a n/a
semiarid
highlands
n/a
Eastern 5 12
temperate
forests
n/a
Marine West 5 n/a
Coast forest
n/a
Northern 5 7
(eastern)
forest
n/a
Northwest 5 10
forested
mountains
n/a
Temperate n/a n/a
sierras
n/a
Great Plains 12 n/a
n/a
Lichens
Mediterranean 3 6
California
n/a
June 2018
6-194
DRAFT: Do Not Cite or Quote

-------
Table 6-28 (Continued): Critical loads for nitrogen by Pardo (2011) and more
recent critical load information.
Pardo etal. (2011c)
New Critical Load Information
Upper
Lower Critical
Critical Load
Ecoregion Load (kg (kg
(Level 1) N/ha/yr) N/ha/yr)
Biological
and
Critical Load Type of Chemical Study
(kg N/ha/yr) Ecosystem Effects Species Reference
North 3 n/a
American
deserts
n/a
Southern n/a n/a
semiarid
highlands
n/a
Eastern 4 8
temperate
forests
n/a
Marine West 2.7 9.2
Coast forest
1.54 and 2.51 Pacific Lichen EDiDhvtic Root et al. (2015)
Northwest communities lichens
and lichen N
concentration
4 Northern Dearadation EDiDhvtic McMurrav et al.
Rockv to lichen lichens (2015)
Mountains communities
Northern 4 6
(eastern)
forests
4-6 Northeastern Decreases in EDiDhvtic Cleavitt et al. (2015)
U.S. Class I species lichens
areas richness and
N sensitive
species, and
poorer thallus
condition
Northwest 1.2 7.1
forested
mountains
<4.1 Wind River Poorerthallus EDiDhvtic McMurrav et al.
Ranae. WY. condition lichens (2013)
including the
Class I
Bridger
Wilderness
Temperate 4 7
sierras
n/a
Great Plains n/a n/a
n/a
Herb and shrub
Mediterranean 6 33
California
<11kaN/ha/vr Semiarid Conversion to Shrubs. Cox et al. (2014)
coastal sage exotic grasses,
scrub, CA grasslands and forbs
June 2018
6-195
DRAFT: Do Not Cite or Quote

-------
Table 6-28 (Continued): Critical loads for nitrogen by Pardo (2011) and more
recent critical load information.
Pardo etal. (2011c)
Ecoregion
(Level 1)
Lower
Critical
Load (kg
N/ha/yr)
Upper
Critical
Load
(kg
N/ha/yr)
New Critical Load Information
Critical Load
(kg N/ha/yr)
Type of
Ecosystem
Biological
and
Chemical
Effects
Study
Species
Reference
North
American
deserts
8.4
Open: 8.3-9.9
(mean = 9.2,
n = 240)
Closed:
13.5-17.0
(mean = 16.5,
n = 32)
Ecoregion
Decreasing
species
richness
Grasses
and forbs
Simkin et al. (2016)
2.1 and 3.6
Southern
California
Joshua Tree
National Park
Exponential
Creosote
increase in
bush
the probability
(Larrea
of biomass
tridentata)
(simulated
other
using the
shrubs,
DayCent
forbs, and
model)
grasses;
exceeding the
two tree
fire threshold
species
of 1,000 kg/ha
included

(California

juniper and

Single leaf

pinyon-

Juniperus

californica

and Pinus

monophylla)
Rao et al. (2010)
Southern
semiarid
highlands
n/a
n/a
n/a
Eastern
temperate
forests
n/a
17.5
Open: 6.6-9.7
(mean = 7.9,
n = 947)
Closed:
7.8-19.3
(mean = 12.5,
n = 7,378)
Ecoregion
Decreasing
species
richness
Grasses
and forbs
Simkin et al. (2016)
June 2018
6-196
DRAFT: Do Not Cite or Quote

-------
Table 6-28 (Continued): Critical loads for nitrogen by Pardo (2011) and more
recent critical load information.
Pardo etal. (2011c)
Ecoregion
(Level 1)
Lower
Critical
Load (kg
N/ha/yr)
Upper
Critical
Load
(kg
N/ha/yr)
New Critical Load Information
Critical Load
(kg N/ha/yr)
Type of
Ecosystem
Biological
and
Chemical
Effects
Study
Species
Reference
Marine West
Coast forest
n/a
n/a
Open: No data
Closed:
10.4-15.0
(mean = 12.8,
n = 24)
Ecoregion
Decreasing
species
richness
Grasses
and forbs
Simkin et al. (2016)
Northern
(eastern)
forests
21
Open: 8.0-9.8 Ecoregion
(mean = 8.9,
n = 75)
Closed:
8.0-18.9
(mean = 13.8,
n = 1,955)
Decreasing
species
richness
Grasses
and forbs
Simkin et al. (2016)
Northwest
forested
mountains
10
Open: 8.0-10.2 Ecoregion
(mean = 9.1,
n = 1,429)
Closed:
10.8-19.6
(mean = 15.3,
n = 2,113)
Decreasing
species
richness
Grasses
and forbs
Simkin et al. (2016)
1.9-3
Subalpine
3.0
Alpine
ForSAFE-
Subalpine
McDonnell et al.
VEG modeled
plant
(2014a)
changes in
community,

subalpine
consisting

plant
primarily of

community
forest

between 2010
understory

and 2100
plants and
one
additional
tree species
(Subalpine
fir-Abies
lasiocarpa)

Protection of
Grasses
Bowman et al.
natural
and forbs
(2012)
community


cover


June 2018
6-197
DRAFT: Do Not Cite or Quote

-------
Table 6-28 (Continued): Critical loads for nitrogen by Pardo (2011) and more
recent critical load information.
Pardo etal. (2011c)
New Critical Load Information
Upper
Lower Critical
Critical Load
Ecoregion Load (kg (kg
(Level 1) N/ha/yr) N/ha/yr)
Biological
and
Critical Load Type of Chemical Study
(kg N/ha/yr) Ecosystem Effects Species Reference
Northwest
forested
mountains
(Continued)
1-2 Abine and ForSAFE- Alpine and Sverdrup et al.
subalpine VEG modeled subalpine (2012)
changes in ground
alpine and vegetation,
subalpine including
plant two tree
community species
(modeled from (Engelmann
1750 to 2500) spruce and
White
spruce-
Picea
engelmannii
and Picea
glauca)
Temperate n/a n/a
sierras
Open: 8.6-8.7 Ecoreaion Decreasina Grasses Simkin et al. (2016)
(mean = 8.65, species and forbs
n = 3) richness
Closed:
14.8-14.8
(mean = 14.8,
n = 42)
Great Plains 5 25
Open: 8.3-9.8 Ecoreaion Decreasina Grasses Simkin et al. (2016)
(mean = 9.3, species and forbs
n = 618) richness
Closed:
11.3-19.6
(mean = 16.6,
n = 274)
Forest
n/a n/a n/a
8 32 forest Saturation of Conifer Fleischer et al.
sites around photosvnthetic (2013)
the globe capacity of the
canopy
Mediterranean 17 39
California
n/a
North n/a n/a
American
deserts
n/a
Southern n/a n/a
semiarid
highlands
n/a
June 2018
6-198
DRAFT: Do Not Cite or Quote

-------
Table 6-28 (Continued): Critical loads for nitrogen by Pardo (2011) and more
recent critical load information.

Pardo etal. (2011c)
New Critical Load Information
Ecoregion
(Level 1)
Lower
Critical
Load (kg
N/ha/yr)
Upper
Critical
Load
(kg
N/ha/yr)
Critical Load
(kg N/ha/yr)
Biological
and
Type of Chemical Study
Ecosystem Effects Species
Reference
Eastern
temperate
forests
3
8
n/a
Marine West
Coast forest
5
n/a
n/a
Northern
(eastern)
forests
3
26
n/a
Northwest
forested
mountains
4
17
n/a
Temperate
sierras
n/a
n/a
n/a
Great Plains
n/a
n/a
n/a
ForSAFE-VEG = a dynamic forest ecosystem model; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
June 2018
6-199
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
APPENDIX 7. AQUATIC BIOGEOCHEMISTRY
This appendix sumarizes recent advancements in understanding the effects of nitrogen
(N) and sulfur (S) deposition on aquatic biogeochemical processes and cycles. Freshwater
systems of the U.S. include lakes (lentic systems), rivers and streams (lotic systems), and
wetlands. The latter, including bogs, fens, marshes, and swamps, are discussed in
Appendix 11 and Appendix 12. The freshwater section (Appendix 7.1) is further
subdivided into N and S sources (Appendix 7.1.1); ecosystem processes, effects, and
indicators (Appendix 7.1.2): monitoring (Appendix 7.1.3); modeling (Appendix 7.1.4);
national scale sensitivity and chemical recovery (Appendix 7.1.5); water quality criteria
(Appendix 7.1.6); and a summary (Appendix 7.1.7).
Appendix 1_1 is an overview of the complex biogeochemical processes affected by N
loading to estuaries (where fresh water from rivers meets the salt water of oceans) and
other near-coastal areas such as lagoons and open ocean areas near coastlines. Inputs of
N, including atmospheric deposition (Appendix 7.2.1). to the highly variable estuarine
environment (Appendix 7.2.2). and factors such as dissolved oxygen (DO)
(Appendix 7.2.3) and pH (Appendix 7.2.4) affect N cycling. Key processes and indicators
of N cycling are discussed in Appendix 7.2.5 and Appendix 7.2.6 followed by monitoring
(Appendix 7.2.7). modeling (Appendix 7.2.8). national scale sensitivity (Appendix 7.2.9).
water quality criteria (Appendix 7.2.10) and a summary (Appendix 7.2.11).
Biological effects and indicators linked to altered aquatic biogeochemistry are discussed
in Appendix 8 (Biological Effects of Freshwater Acidification), Appendix 9 (Biological
Effects of Freshwater Nitrogen Enrichment), Appendix 10 (Biological Effects of
Nitrogen Enrichment in Estuaries and Near-Coastal Systems), and Appendix 12
(Non-Acidifying Effects of Sulfur).
7.1. Biogeochemistry of Nitrogen and Sulfur in Freshwater
Systems
As described in the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria (hereafter referred to as the 2008 ISA), the most common and
well-documented aquatic effects are acidification and eutrophication, which may occur
simultaneously in some water bodies. Since the 2008 ISA, there have been additional
estimates of the proportion of total N loading in freshwater systems attributed to
atmospheric deposition and refinements in spatial and temporal trends of N and S
deposition across the U.S. Chemical indicators of N deposition identified by the 2008
June 2018
7-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
ISA were nitrate (NOs ) and dissolved inorganic nitrogen (DIN) concentrations in surface
waters. Surface water chemistry indicative of acidic conditions and acidification effects
includes concentrations of sulfate (SOr ). NOs . inorganic aluminum (Al), calcium (Ca),
sum and surplus of base cations, acid neutralizing capacity (ANC), and surface water pH.
Continued monitoring of these indicators provides some evidence of chemical recovery
from acidification in some U.S. surface waters. A number of studies since 2008 have
focused on improving understanding of aquatic acidification and eutrophication processes
mediated by N. Many of these have focused on pathways of pollutant and other
constituent movement within ecosystems, including monitoring studies of various kinds.
The body of evidence is sufficient to infer a causal relationship between N and S
deposition and the alteration of freshwater biogeochemistry, which is consistent with
the conclusions of the 2008 ISA.
Acidification of freshwater ecosystems occurs in response to either S or N deposition
alone or in combination. This is because both N and S deposition can act as acidifying
agents. Acidifying deposition effects on biogeochemical processes in soils (Appendix 4)
have significant ramifications for the water chemistry and biological functioning of
associated surface waters. Surface water chemistry integrates direct air-to-water
deposition with deposition impacts on soil chemistry of hydrologically connected
terrestrial ecosystems within the watershed. Deposited N and S interact with the soils and
sediments of terrestrial and aquatic ecosystems via oxidation and reduction reactions as
well as biological uptake and microbially mediated processes.
Acidification of fresh water may occur as a chronic or an episodic condition
(Appendix 8.2). The 2008 ISA defined chronic acidification by reference to annual
average conditions, often quantified as summer and fall chemistry for lakes and as spring
baseflow chemistry for streams. Episodic acidification refers to conditions during
rainstorms or snowmelt when water has a relatively short residence time in soil before
entering surface water. During these stormflow events, proportionately more water drains
through upper soil horizons, providing less neutralization of atmospheric acidity
compared with flow through deeper soil horizons. Episodes of acidification are typified
by lower pH, lower ANC, and higher inorganic Al concentration in surface water than
during baseflow conditions. The short-term change in chemical conditions may be toxic
or lethal to aquatic biota (Appendix 8). These conditions are caused partly by acidifying
deposition (SO42 , NO3 ) and partly by natural processes, including base cation dilution
and flushing of organic acids into drainage water. It is known that the biota in many
streams/lakes are impacted when the ANC is consistently below 50 ueq/L. For this
reason, the U.S. EPA National Lakes Assessment used an ANC threshold of >50 ueq/L as
indicative of nonacidified water bodies (U.S. EPA. 200%).
7-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
In aquatic systems, N is a nutrient that stimulates growth of primary producers (algae
and/or aquatic plants). Even small inputs of N in low nutrient water bodies such as
remote headwater and lower order streams and alpine lakes can increase nutrient
availability, alter the balance of N and phosphorus (P) nutrients, affect biogeochemical
processing of N and increase the productivity of photosynthesizing organisms, resulting
in a increase in pool of fixed carbon (C). Nutrient enrichment leads to changes in aquatic
assemblages and biodiversity in freshwater (Appendix 9) and coastal regions
(Appendix 10). The freshwater ecosystems in the U.S. most likely to be sensitive to
nutrient enrichment from N deposition are headwater streams, lower order streams, and
alpine lakes which have very low nutrients and productivity and are far from local
pollution sources [Appendix 9.1.1.2;(U.S. EPA. 2008a)l. High-elevation lakes in the
western U.S. are naturally oligotrophic and are considered among the aquatic ecosystems
most sensitive to N deposition (Williams et al.. 2017b). A portion of these lakes and
streams in the western U.S. are in Class I wilderness areas (Williams et al.. 2017b; Clow
et al.. 2015; Nanus et al.. 2012). In higher order streams, N deposition typically mixes
with N derived from other nonatmospheric sources, including urban/suburban point and
nonpoint sources, industrial sources, and agricultural sources. The long-held paradigm
that fresh waters are typically P limited has been replaced by an understanding that both
N and P can limit primary production, and that nutrient limitation can be a dynamic and
transient process (Paerl et al.. 2016b; Paerl et al.. 2014; Conlev et al.. 2009; Paerl. 2009).
Although over-enrichment with P is not specifically addressed in this assessment, recent
trends in increased atmospheric deposition of P (Appendix 9.1.1.2) have ramifications for
nutrient stoichiometry.
The geochemical processes and associated chemical indicators discussed in this appendix
can be considered to indicate or suggest eutrophication or acidification. Some of the
biogeochemical alterations associated with N and S deposition link directly to the
biological effects discussed in subsequent appendices. Others do not cause direct
biological effects but are precursory steps to changes in soil or water chemistry that can
lead to biological effects. Table 7-1 summarizes the key freshwater indicators for N
driven nutrient enrichment and acidification and the section of the ISA that discusses
each endpoint.
A separate appendix addresses the biogeochemistry of terrestrial responses to nutrient
and acidic additions (Appendix 4). Aquatic and terrestrial systems are interconnected and
many of the biogeochemical processes discussed herein bridge the transitions between
these two environmental compartments.
7-3
DRAFT: Do Not Cite or Quote

-------
Table 7-1 Summary of key freshwater indicators of eutrophication and
acidification.
Endpoint
N Driven Nutrient
Enrichment
Acidification
Section of ISA That
Discusses Endpoint
Chemical indicator
Water NO3" concentration
X
X
7.1.2.1
Water SO42" concentration

X
7.1.2.2
Water pH

X
7.1.2.5
Water ANC

X
7.1.2.6
Water base cation surplus

X
7.1.2.7
Water inorganic Al concentration

X
7.1.2.8
Biological indicator
Diatoms
X

9.2.1
Nutrient ratios
X

9.2.2
Phytoplankton biomass shift
X

9.2.3
Periphyton/microbial biomass
X

9.2.4
Chlorophyll a
X

9.2.5
7.1.1.	Nitrogen and Sulfur Sources
1	In fresh water, both S and N can contribute to acidification while nutrient enrichment
2	effects are associated with N. Long-range atmospheric transport of N and S can affect
3	remote freshwater catchments far from pollutant sources. Major sources of N and S and
4	deposition trends are discussed in Appendix 2.
7.1.1.1.	Nitrogen Sources
5	Since the 2008 ISA, additional analyses have refined the understanding of N sources and
6	deposition trends (Appendix 2) to freshwater systems. Sobota et al. (2013) synthesized
7	data on N inputs to lands and waterways throughout the U.S. They found that
8	human-caused N inputs are ubiquitous, but are spatially heterogeneous. The highest N
June 2018	7-4 DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
loads generally occurred in the Midwest, Mid-Atlantic region, central and southern
California, and portions of the Columbia River Valley. Synthetic fertilizer was estimated
to be the single largest source of human-caused N inputs to 41% of the water resource
units analyzed, followed by atmospheric deposition (33%) and biological N fixation
(22%). A tool for calculating net anthropogenic nitrogen inputs (NANI) was developed
for watersheds across the contiguous U.S. at the county level (Hong et al.. 2013. 2011).
The NANI Calculator Toolbox takes into consideration fertilizer N application,
agricultural N fixation, net food and feed imports, and atmospheric sources of N. Unlike
S, geological sources of N are rare.
In a U.S. Geological Survey (USGS) National Water Quality Assessment (NAWQA)
report on occurrence and distribution of nutrients in streams and groundwater (based on
water quality assessments conducted from 1992 to 2001), atmospheric deposition was
identified as the largest nonpoint source of N in the less developed watersheds (areas
dominated by forest or rangeland with <5% urban and <25% agricultural land) in the
eastern part of the country where deposition rates have historically been highest, in areas
near the Great Lakes, and the mountainous west (Dubrovskv et al.. 2010). Atmospheric
sources have been shown to be quantitatively important (>33% of total input) to Lake
Tahoe, CA/NV (Sahoo et al.. 2013; Dolislager et al.. 2012). Flathead Lake, MT (Ellis et
al.. 2015). and Nine Mile Run in Pittsburgh, PA (Divers et al.. 2014). Even in lowland
lakes, N from atmospheric sources has been shown to contribute appreciably to the total
input of N. In a study of nutrient sources to Saginaw Bay, MI, N deposition was
estimated to comprise 10 to 11% of the total N input from 1987 to 2002 (He et al.. 2014).
Nitrogen is deposited in various reduced and oxidized forms, including organic N, and, in
wet or dry forms as well (Appendix 2). Oxidized N is emitted into the atmosphere mainly
from motor vehicles, electricity generating units, and industry. Reduced N is emitted
mainly from agricultural sources such as livestock and fertilizer applications. Direct
deposition of N to open water surfaces constitutes an appreciable source of N to relatively
large freshwater lakes and rivers, estuaries, and coastal marine waters (Appendix 7.2.1).
Alternatively, N can be deposited in the watershed of receiving waters and then move
through soils before entering surface waters. Atmospheric deposition of reduced N has
increased relative to oxidized N in parts of the U.S. including the East and Midwest in the
last few decades, shifting from a NOs dominated to an ammonium (NH4 ) dominated
condition, and this trend is expected to continue under existing emissions controls (Li et
al.. 2016d; Pinder et al.. 2008; U.S. EPA. 2008a). Since the 2008 ISA, there is a greater
understanding of the potential ecological changes associated with the increase in
deposition of reduced N relative to oxidized N in terrestrial, freshwater and coastal
systems (Chapter 1.2.2.5V
7-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
Table 7-2 summarizes key studies that have quantified total N loading in freshwater
systems from atmospheric deposition. In the 2008 ISA, the difficulty in determining the
percentage of atmospheric N in lowland waters was noted because there are so many
other point and nonpoint sources of N to drainage waters in these areas (U.S. EPA.
2008a). As described in the 2008 ISA, a large fraction of atmospheric N deposition is
retained in most forested ecosystems with less retention in urban, suburban, and
agricultural lands (U.S. EPA. 2008a). The atmospheric fraction that does leach to streams
can make a substantial contribution to total N inputs to downstream waters, especially in
the eastern U.S. (Driscoll etal.. 2003d). Apportionment of N sources to identify the
contribution of atmospherically derived N was described in a few studies in the 2008 ISA
(U.S. EPA. 2008a). In Spatially Referenced Regressions on Watershed Attributes
(SPARROW) modeling studies, approximately 70% of the N in headwater streams in the
northeastern U.S. was judged to be from N deposition (Alexander et al.. 2007; Alexander
et al.. 2002). This SPARROW model application used only wet deposition as the measure
of N deposition. A large amount of N from nonpoint source urban influences (this is most
likely primarily the dry deposition of exhaust N gases) often approximately doubles the
importance of N deposition as an N source to higher-order river systems (Howarth.
2008a. b).
Several new studies published since the 2008 ISA further quantified N sources, including
atmospheric contribution, to lakes and streams (Table 7-2). Sebestven et al. (2008)
showed the dominant direct role of atmospheric NOs in snowmelt runoff at Sleepers
River, VT. At peakflow, 48% of NO3 was from atmospheric sources in this forested
watershed. In the Uinta Mountains, UT, at least 70% of NO3 was atmospherically
deposited to high alpine lakes from distant anthropogenic sources (Hundev et al.. 2016).
Approximately 60% of the reactive N was from agriculture, based on isotopic analysis
and modeling of snow, inflow and lake NO3 . The authors suggested these findings are
widely applicable to other western U.S. alpine sites based on comparison of isotope and
precipitation data in the region.
7-6
DRAFT: Do Not Cite or Quote

-------
Table 7-2 Summary of recent studies quantifying nitrogen deposition
contribution to total nitrogen loading in freshwater systems.
Region
Total N Loading Due to Atmospheric
Deposition
Method
Reference
Lake Tahoe, CA Approximately 57%, NH3 as the dominant
component and nitric acid and NO3" representing
a smaller but not insignificant proportion of total
N.
Field data and
Lake Tahoe
watershed model
Sahoo etal. (2013)
Dolislaqer et al.
(2012)
Flathead Lake, MT
Atmospheric loading of NhU"1" averaged 44% of
total load between 1985 and 2004 and was the
primary form of N in deposition.
Field data and
statistical analysis,
linear regression
Ellis etal. (2013)
Unita Mountains, UT At least 70% of NO3" in the high alpine lakes is
from atmospheric deposition. Most reactive N
originates from agricultural activities
(approximately 60%).
Isotopic analysis
and baysian-based
stable isotope
mixing model
Hundev etal. (2016)
Saginaw Bay, Ml
N deposition was estimated to be 10 to 11% of
total N from 1987-2002.
Multiple databases
of land use/cover,
hydrography,
animal production,
fertilizers,
combined
wastewater
overflows
He et al. (2014)
St. Lawrence River,
Quebec City
Atmospheric N contributed about 0 to 4% of total
N load in the summer; in the spring, this source
represented 4 to 11% of total N.
Isotopic analysis Thibodeau et al.
(2013)
Quinnipiac River,
CT
Atmospherically deposited NO3" represented
<6% of N loading; however, during storm events,
atmospheric deposition represented up to 50% of
stream NO3" but varied widely by site.
Stable isotope
ratios
Anisfeld et al. (2007)
Baltimore Long- Atmospheric contributions ranged from 5 to 94%
Term Ecological during storm flow conditions and represented
Research site	approximately 50% of the highest NO3" loads
during storms.
N watershed mass
balances and
stable isotopes
Kaushal etal. (2011)
Tributary to Teff Contributions of atmospheric N to surface water
Run, MD	in an Appalachian Mountain stream were most
evident during hydrological episodes, but
baseflow accounted for much of the NO3" loss in
stream water.
Isotopic analysis Sabo et al. (2016)
Nine Mile Run in 34% of NO3" in stream water was atmospheric in
Pittsburgh, PA	origin during storm events while 94% of stream
water NO3" was from sewage sources during
baseflow conditions.
Stable isotope
ratios
Divers et al. (2014)
Suburban
watershed (Lisha
Kill) in eastern New
York
About 40% or more of NO3" in stream water in a Stable isotope
suburban watershed during storm runoff was ratios
attributed to direct NO3" deposition.
Burns et al. (2009)
June 2018
7-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
Table 7-2 (Continued): Summary of recent studies quantifying nitrogen deposition
contribution to total nitrogen loading in freshwater
systems.
Total N Loading Due to Atmospheric
Region	Deposition	Method	Reference
Sleepers River
In this forested watershed 48% of NO3" was from
Isotopic tracers
Sebestven et al.
watershed,
atmospheric sources at peakflow. More than half
and mixing
(2008)
northeastern
of the NO3" in soil and shallow groundwater after
analysis

Vermont
the start of snowmelt originated directly from
atmospheric deposition.
Stable isotope
ratios,
hydrochemistry,
end-member
mixing analyses

Key pre-2008 literature
16 northeastern Approximately 70% of the N in headwater	SPARROW
watersheds	streams was from N deposition and the net
transport of N from headwater streams was
between 40 and 65% of total N.
N deposition contributes approximately 31% of	Bover et al. (2002)
the total N load to large river basins, although this
fraction varies regionally. Values for watersheds
in northern New England were substantially
higher and atmospheric deposition dominated.
Eight watersheds in Atmospheric deposition was second largest N PnET-BGC and Driscoll et al. (2003d)
New York and New input for the eight watersheds (11 to 36% of total) WATERS N
England	with four watersheds ranging from 34 to 36%.
N = nitrogen; NH3 = ammonia; NH4+ = ammonium; N03 = nitrate; SPARROW = Spatially Referenced Regressions on Watershed
Attributes.
Additional information is available since the 2008 ISA on atmospheric contributions of
reduced versus oxidized N to lakes. In Lake Tahoe, where a total maximum daily load
(TMDL) allocation was adopted in 2011 by California and Nevada to reduce nutrient
loading, atmospheric N represented approximately 57% of the total N loading to the lake
(Sahoo et al.. 2013). In this large alpine lake in the Sierra Nevada mountain range, an
estimated 185 metric tons of N was directly deposited from the atmosphere, with NH4 as
the dominant component and nitric acid plus NOs representing a smaller, but not
insignificant, proportion of total N (Dolislager et al.. 2012). In Flathead Lake, MT,
atmospheric loading of NH44" averaged 44% of the total load between 1985 and 2004 and
was the primary form of N in atmospheric deposition (Ellis et al.. 2013). There was an
increase in atmospheric loading of NO3 + nitrite (NO2 : +48%) and NH4 (+198%) and
decrease in total P loading (-135%).
In the studies reviewed in the 2008 ISA, the role of atmospheric deposition in
downstream urban and residential water bodies was rarely addressed (U.S. EPA. 2008a).
Alexander et al.
(2002):
Alexander et al.
(2007)
16 large
northeastern river
basins
June 2018
7-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
New studies using stable isotope ratios to quantify N loading in streams have
characterized the contributions of atmospheric N in these systems. Several of these
studies have shown shifts to higher atmospheric N contributions during storm events. In
the Quinnipiac River in Connecticut which drains into Long Island Sound,
atmospherically deposited NOs averaged <6% of average N loading during baseflow
conditions; however, during storm events, atmospheric deposition represented up to 50%
of stream NO3 . although the amount varied widely by site (Anisfcld et al.. 2007). In the
St. Lawrence River outlet at Quebec City, Canada, isotopic analysis of N inputs indicated
that atmospheric N contributed about 0 to 4% of total N load in the summer, while in the
spring this source was 4 to 11% of total N (Thibodcau et al.. 2013). In forested,
agricultural and urban watersheds at the Baltimore Long-Term Ecological Research site,
atmospheric N contributions ranged from 5 to 94% during storm flow conditions and
represented approximately 50% of peak storm NO3 (Kaushal et al.. 2011). In Nine Mile
Run in Pittsburgh, PA, 34% of NO3 in stream water was atmospheric in origin during
storm events, whereas during baseflow conditions 94% of stream water NO3 was from
storm drain and wastewater treatment sources (Divers et al.. 2014). In another study in
Pennsylvania, stream water from Spring Creek was monitored at sites from upstream to
downstream during storm flow events to assess changes in N sources (Buda and
DeWalle. 2009). For the forested upstream site, the atmospheric contribution varied by
storm size. In the downstream urbanized watershed, atmospheric NO3 was an important
N source at peak flow, especially during short-duration storms when overland flow was
prevalent.
Burns et al. (2009) applied dual isotope analysis of NO3 to determine the dominant
sources and processes that affect NO3 concentrations in six streams on different land
uses in New York. The dual isotope data revealed varying sources and processes that
affect N03 concentration among the six stream watersheds. Atmospheric NO3 was
about half of stream NO3 during storms because of rapid shunting of runoff through
storm drains. The suburban watershed that had no septic or wastewater influence showed
N03 concentrations only slightly higher than those observed in two forested watersheds.
Overall, these studies appear to have characterized relatively well the atmospheric
contribution of N to water bodies that have multiple sources of N. Results show
substantial variability, with atmospheric sources typically being most pronounced during
high-flow conditions.
7.1.1.2. Sulfur Sources
The 2008 ISA reported that there were both depositional and geological sources of SO42
to aquatic ecosystems. SOx deposition to ecosystems is primarily in the chemical form
June 2018
7-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
SOr . which is a moderately mobile anion in the soil solution and surface water of many
glaciated acid-sensitive watersheds in the northeastern U.S. Note that SOr is less mobile
in unglaciated watersheds of the southeastern U.S. (Rice et al.. 2014). Other major
sources of SOr in ecosystems include mineralization of S from organic matter and
weathering of geologic sources of S, including pyrite minerals. Anthropogenic soil and
rock disturbance in the form of road cuts or mine tailings can increase geological
contributions of SO42 to soil or surface water by exposing S bearing minerals to oxygen
(O2), promoting mineral oxidation and the release of SO42 . Acid mine drainage is one
result of this process, and is prevalent in many parts of the Appalachian Mountain region.
New published evidence regarding the influence of geologic sources of S or acid mine
drainage on stream chemistry has not been reviewed for this ISA.
The 2008 ISA described how SOx deposition causes the release of S from terrestrial to
aquatic ecosystems via SO42 leaching, and other changes in surface water chemistry
caused by increasing SO42 concentrations. At that time, acidification of surface waters at
most locations in the U.S. where acidification has been documented in response to
deposition was caused mainly by SO42 (Driscoll et al.. 2001b; Sullivan. 2000). This was
partly due to the mobility of SO42 in many (especially northeastern) waters draining
from terrestrial ecosystems to aquatic ecosystems. It was also known that the mobility of
S042 varies geographically, with S adsorption on soils most significant in regions that
were unglaciated during the most recent glaciation. This includes many acid-sensitive
watersheds in the southeastern U.S. In the most recently glaciated Northeast and upper
Midwest, a substantial component of the atmospherically deposited S accumulates in
organic pools in soil, although microbial mineralization can transform this organic S back
into more mobile SO42 .
A mass balance study by Mitchell et al. (2011) of 15 watersheds located in the
northeastern U.S. and southeastern Canada suggested substantial sources of SO42 in
watershed soils. The internal S sources were attributed mainly to mineralization of S
stored in soils in response to decades of atmospheric S deposition. In a study of 16 lakes
from the original Adirondack Long-Term Monitoring program from 1984 to 2010
Mitchell et al. (2013) observed discrepancies in the calculated S mass balances that were
associated with discharge. These results suggested that internal S sources have become
more important to the watershed S budget as atmospheric S deposition has decreased.
Declines in lake SO42 concentrations have been observed in locations where S deposition
has decreased significantly, such as in the Adirondack Mountains KMitchcll et al.. 2013);
Appendix 7.1.5.11.
7-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
7.1.2.
Ecosystem Processes, Effects, and Indicators
New information on biogeochemical indicators and processes of eutrophication and
acidification are presented in the following sections, along with summaries of
information from the 2008 ISA. These processes may occur either in sequence or
simultaneously in a given geographic area. Impacts of N and S deposition on aquatic
ecosystems can be described by changes in acid-base chemical indicators including SOr
concentration, NOs concentration, DIN concentration, inorganic A1 concentration, base
cation concentrations, base cation surplus (BCS), pH, and ANC. Surface water NO3
concentration can reflect both eutrophication and acidification. Water pH and the
concentrations of ANC, SO42 . and inorganic A1 are commonly used indicators of the
likelihood of surface water acidification (U.S. EPA. 2008a). As reported in the 2008 ISA
and summarized in Figure 7-1. the biogeochemical cycles of N, P, and C are linked in
freshwater ecosystems. Nitrogen deposition can alter the pools and fluxes of the C, N,
and P cycles, including nitrification and denitrification. Sulfur deposition adds SO42
directly to soil solutions and to surface waters with effects on ecosystems (Appendix 12).
but many ecological effects are mediated through the indirect effects of SO42 on the
exchange of acidic and basic cations on soils. The chemical indicators of deposition
discussed below also link to biological effects of acidifying deposition (Appendix 8) and
freshwater N enrichment (Appendix 9).
June 2018
7-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
CO.
Deposition
Terrestrial Input
CCh
CO,
Deposition
[ *
CO,
s
1
to
EC
LJp1ak<2
				 - ¦§
ExcrtlloA
*
Primary t
? Producers

Uptake
En.c*tK5n

%
%
-i:
V
CYIC
ANOXIC
GfazEfs,
Fredalflra,
and Viruses
i Ł
7 ""H

ft


r
~ i

c
WVNH.*

Ł
0
1—

c.
!
CU

31
~

0 |


H 1


li
i

•[ Xfi
Detritus
OxompojltKan j
NOj-JNCV
Ssdimsnls
Btrvihlc
Oecompc-Eiton
C = carbon; C02 = carbon dioxide; N = nitrogen; P = phosphorus.
Modified from U.S. EPA (2008a).
Figure 7-1 Nitrogen cycle in freshwater ecosystem.
7.1.2.1. Nitrogen in Surface Waters
As reviewed in the 2008 ISA and supported by more recent studies, the fate and transport
of deposited N is influenced by characteristics of the catchment and the receiving waters.
Retention of N varies among watershed types such that similar amounts of deposition can
result in different rates of N leaching, depending on catchment characteristics
[Appendix 9.1.1.2;(Bergstrom. 2010)1. In most surface waters of the U.S., dissolved
inorganic N (DIN, the sum of the concentrations of NO3 , NH4+, and nitrite) is
overwhelmingly dominated by NO3 . As summarized in the 2008 ISA, high
concentrations of NO, in lakes and streams, indicative of terrestrial ecosystem N
saturation, have been found at a variety of locations throughout the U.S. Surface water
NO;, is a chemical indicator for both eutrophication and acidification. Nitrate contributes
June 2018
7-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
to the acidity of many lakes and streams in the eastern U.S. that have been affected by
acidifying deposition, especially during spring snowmelts and under high-flow
conditions.
In the 2008 ISA, a study evaluating the relationship between wet deposition and DIN
concentration in 4,296 lakes across Canada, Europe, and the U.S. showed a significant
correlation between increases in N wet deposition and increases in lake DIN
concentration (Bcrgstrom and Jansson. 2006). More recent studies from some regions of
the U.S. [(e.g., Driscoll et al.. 2016; Eshleman and Sabo. 2016; Strock et al.. 2014;
Eshleman ct al.. 2013); Appendix 7.1.5.11 showed recent declines in concentrations of
N03 in surface waters that are consistent with declines in N deposition. Using the Lake
Multi-Scaled Geospatial and Temporal Database of the Northeast Lakes of the U.S.
(LAGOS-NE) containing water quality data from 2,913 lakes, Oliver et al. (2017)
identified atmospheric deposition as the main driver of declines in total N (TN)
deposition and lake TN:total P (TP) ratios from 1990 to 2011.
Reactive N can be taken up by terrestrial and aquatic biota or stored in soils and/or
sediments, be re-emitted back to the atmosphere by the microbial process of
denitrification (Appendix 7.1.2.3). or carried downstream in a dissolved state, most
typically as NO3 , because much of the deposited NH4 is either taken up by biota or
nitrified to NO3 . Typically, a rather large percentage of the N deposition to a given
watershed is taken up or stored and is not denitrified or made available for leaching to
surface waters. On average, for a large watershed with mixed land use, it has been
estimated that roughly 75% of N inputs are retained in the watershed and 25% are
exported downstream regardless of the dominant N input, including deposition (Howarth
et al.. 2012; Howarth et al.. 1996a). This stored N does not contribute directly to water
acidification or eutrophication and represents a significant portion of the incoming N
deposition to forested ecosystems in the U.S. The portion of the N deposition that is not
retained or lost to denitrification can contribute to water acidification or eutrophication,
both on a chronic and especially an episodic basis.
It was well understood at the time of the 2008 ISA that streams and their associated
sediments can transform nutrients, store them for the short term, or serve as a sink for N
loss from the watershed through denitrification. It is important to consider the balance
between how much and how long streams retain elements versus transport them
downstream. This balance is key to modeling watershed nutrient export. Headwater
streams can play disproportionately large roles in N transformation and N cycling in
aquatic ecosystems, and they are important to predicting effects of N loading on
downstream ecosystems. Atmospheric deposition may represent a significant source of N
to some headwater streams r(Lawrence et al.. 2015b); Table 7-21. which often have
7-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
increased water residence time and solute retention due to increased interactions between
surface and groundwater in the hyporheic zone.
The majority of atmospherically deposited N is either denitrified or accumulates in
watershed soils, vegetation, or groundwater (Galloway et al.. 2008). The relative
partitioning of N loss via denitrification versus watershed storage is poorly known
(Galloway et al.. 2004). Watershed denitrification hot spots and hot-moments tend to
occur in areas characterized by wet conditions, including wetlands and riparian zones.
Such ecosystems can be highly efficient in removing N from surface waters via
denitrification (Galloway et al.. 2003). Nitrogen in fresh waters is also controlled by
hydrological conditions. For example, Lusardi et al. (2016) showed that spring-fed rivers
of northern California had higher NOs and PO4 concentrations and cooler temperatures
than rivers that were fed more by runoff. Recent studies have added more quantitative
context to hydrologic processes that control nutrient cycling, including hydrologic
exchange between the streams and the groundwater. This hyporheic exchange can
contribute to N retention in streams because it promotes denitrification associated with
anoxic flow paths through organic stream bottom substrates. For example, Hall et al.
(2009b) observed in a tracer study in the Sawtooth Mountains of Idaho that assimilation
and hydrologic storage can be important for retaining N at the watershed scale. The study
stream was not solely a conduit for nutrients at high flow. It had as high an uptake
velocity for N during the snowmelt high flow as during summer baseflow.
Influence of beaver and human-made dams on NO3 uptake in a small headwater stream
in the Rocky Mountains was investigated by Hubbard et al. (2010). Dams influenced
stream water geochemistry because a portion of the stream water flow was diverted into
the more reactive hyporheic zone. Highly oxic and anoxic regions of the streambed
developed around the more permanent structures. This enhanced N cycling in the
hyporheic zone and increased the potential for denitrification and NO3 uptake. In parts of
the western U.S. Rocky Mountains, there was a strong positive correlation between
surface water NO3 concentration and atmospheric N deposition (Elser et al.. 2009b;
Bergstrom and Jansson. 2006). Such a correlation does not always occur elsewhere
because nonatmospheric watershed sources ofN are often larger than atmospheric
sources, depending mainly on land use. Analysis of increasing NO3 trends over three
decades in streams draining into Lake Ontario, Canada, Eimers and Watmough (2016)
suggested that tributary loading, rather than atmospheric deposition, may have played a
dominant role in causing the recent observed increase in NO3 concentrations in the
waters of Lake Ontario. The authors pointed to an observed trend in Ontario towards
more annual grain crop production that required more N fertilizer addition and less
reliance on perennial crops. This shift in crops could be contributing to the observed
increase in NO3 concentration in the lake.
7-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Studies have been conducted in recent years to elucidate processes affecting surface
water NO3 concentrations, including experimental studies, isotopic analyses, monitoring,
and observational studies. The concentration of NOs may vary widely between baseflow
and peakflow conditions (Table 7-2). Seasonality in surface water chemistry can be
caused, in part, by differential hydrologic flows from watershed seeps and glaciers. These
flows can influence NOs leaching to surface waters. O'Driscoll and DeWalle (2010)
showed that seeps are generally NO3 sinks with NO3 concentrations decreasing
downslope from seep locations in Baldwin Creek in southwestern Pennsylvania. During
cold and wet periods, however, the seeps frequently acted as NO3 sources to the stream.
Locations where upwelling groundwater saturates the surface for most of the year and
excess groundwater can be delivered to the stream channel via surface flow paths may
provide seasonally linked water quality functions that can modify the effects of N
deposition.
Changes in the amount of NO3 leaching and disruption of N cycling can occur with
events in the watershed, including wildfire and timber extraction. Human activities have
contributed to greater frequency and magnitude of wildfire in many parts of the U.S.,
which is well known to increase NO3 leaching. The area impacted by fire has increased
in response to climate change, past fire suppression, and increased human presence in
forested ecosystems (Bladon et al.. 2014). Extreme events have the potential to delay
ecosystem recovery from past water acidification. Evans etal. (2017) assessed the role of
wildfire in regulating the chemistry of a mountain lake in a moorland catchment in the
U.K. affected by a large fire. The lake had been subjected to more than two decades of
ecosystem monitoring prior to the fire. The most pronounced change in lake chemistry
was a substantial increase in lake NO3 concentration to peak values of 111 j^icq/L
(1.55 mg/L) 2 years afterthe fire. PastN deposition may have been high enough to make
the catchment more susceptible to NO3 leaching and accompanying reacidification in
response to the fire (Figure 7-2). The increase in the concentration of this mobile acid
anion increased lake acidity and the concentration of inorganic Al. These changes
represented a reversal in the ongoing documented recovery from past acidification. Forest
disturbance, including clear-cut logging, also contributes to increased leaching of
inorganic N, with potential impacts on downstream nutrient concentrations and acid-base
chemistry. Schelker et al. (2016) evaluated N dynamics in Sweden over a 10-year period
after tree harvesting. DIN leaching in first order streams increased substantially
(-15 times) subsequent to logging. In the larger streams, NO3 leaching was seasonal and
increased in response to logging in the mid-sized, but not the largest, streams. Thus, the
increased mobilization of NO3 in first-order streams caused by logging can affect some
downstream locations at some times.
7-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Nitrogen deposition to snow and glaciers is an important source of N to alpine lakes and
streams which are fed by meltwaters. Since the 2008 ISA, several studies indicate that
glacial meltwater has higher NOs than snowmelt water in some regions of the U.S. This
may influence interpretation of biological data from high elevation lakes and streams
KSlcmmons et al.. 2015; Slemmons et al.. 2013; Saros et al.. 2010; Baron et al.. 2009);
Appendix 9.3.2.11. In two sets of high-elevation lakes in the U.S. Rocky
Mountains—those fed by snowpack melt alone and those fed by both glacial and
snowpack meltwaters—the NO3 concentrations in the glacially influenced lakes were
one to two orders of magnitude higher than in the lakes that were fed only by snowmelt
(Saros et al.. 2010). The higher concentration of NO3 in glacial meltwater relative to
seasonal snowpack meltwater was attributed, at least in part, to reduced contact with
watershed soils where microbial communities could rapidly assimilate the available N.
This contrasts with snowmelt watersheds where the meltwater typically percolates
through soils before reaching the streams (Saros et al.. 2010). Williams et al. (2007)
found that, in the Colorado Front Range, rock/glacier outflow had NO3 concentrations of
69 |icq/L compared to snow with 7 (j,eq/L. In two proximal lakes in the central Rocky
Mountains, NO3 concentration from glacier-fed Jasper Lake, was 2 (j,eq/L compared to
the snowpack-fed Lake Albino where NO3 concentration was only 0.03 (j,eq/L
(Slemmons et al.. 2015). In contrast, in the Northern Cascade Mountains, WA,
differences in NO3 concentration between glacier-fed and snowpack-fed lakes were
much smaller and not statistically significant in some years (Williams et al.. 2016b).
7-16
DRAFT: Do Not Cite or Quote

-------
Nitrate
Sulphate
f ^ $ 4? Ł ^ Ł $ $ rf ^ $
^ ^./ / /• $ # Ł rf # $ 4? /
Inorganic aluminium
# $¦ ^ / .# # Ł ^ ^ -p? # 4
ANC
N# /- / $ # # # $ # 4? 4> 4!
Calcium
E 40
./ ,#¦^./ ^ #
DOC

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
Advances in isotopic analyses have improved techniques for investigating the role of
N03 in N cycling. Curtis et al. (2012) used the dual isotope technique at four moorland
watersheds in Great Britain to investigate NOs production in surface waters. An
estimated 79-98% of the annual median NO3 had been microbially produced, indicating
that both reduced and oxidized N deposition may cycle through the microbial flora and
contribute similarly to NO3 leaching. This is important because atmospheric deposition
of NH4+ has been increasing in many areas of the eastern U.S. while deposition of
oxidized N forms (NOy) has decreased (Appendix 2). Goodale et al. (2009) characterized
the amount, form, isotopic composition (15N and 180), and seasonality of stream N in
forested subwatersheds of the Susquehanna River. Atmospheric deposition contributed
substantial N to this river, which provides two-thirds of the annual N load to Chesapeake
Bay. Retention of atmospherically deposited N in the watershed was estimated to be
higher than 95%. Nevertheless, stream NO3 patterns exhibited substantial seasonality.
Peak values of NO3 (14-96 |icq/L) were reached during summer. Lowest values
(<1 (j,eq/L) occurred in October. The summer's increase in net soil nitrification and
in-stream heterotrophic N uptake in response to litterfall during autumn were identified as
likely important drivers of the observed NO3 seasonality. Other studies that included
mixed land uses found far lower N retention at larger spatial scales, including the
Susquehanna River (Howarth et al.. 2006; Bover et al.. 2002). Using stable isotope
analysis, atmospheric deposition was identified as an important source of stream NO3
concentration at a low NO3 site at Fernow Experimental Forest, WV (Rose et al..
2015b). Based on more than 30 years of monitoring data, three hardwood watersheds
appeared to be less responsive to changes in N deposition than the one study watershed
that had coniferous vegetation. The percentage of stream NO3 concentration contributed
by atmospheric deposition in the study watersheds increased during high-flow periods.
Watershed N budgets and empirical models can be useful for assessing the relative
magnitude and sources of N inputs and losses to watersheds via riverine export. A variety
of computational approaches can be used. Alexander et al. (2002) compared predictions
of total N and stream N transport of six watershed models and showed that models with
greater precision were those that had the most detailed descriptions of N sources, water N
attenuation, and water flow paths. Han and Allan (2008) compared N budget estimation
approaches for 18 catchments located in the vicinity of Lake Michigan. The most robust
model suggested that riverine N export constituted about 21% of N inputs. Using a larger
data set, Howarth et al. (2012) estimated that 25% of N inputs are exported from the
watershed on average. The improved ability to predict N export using methods that
incorporated description of agricultural N sources was particularly noticeable when
applied to watersheds that had pronounced diversity of land use and geology (Han and
Allan. 2008). Comparisons of watershed N budgets across the U.S. indicates that there is
geographic variation in the proportion of N inputs that are exported, and the scaling of
7-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
export with runoff indicates that drier conditions may lead to lower proportions of N
inputs that are exported by rivcrs(Sobota et al.. 2009).
Overall, some studies show recent declines in concentrations of NOs and TN in surface
waters that are consistent with declines in N deposition. Advances in isotopic analysis
have allowed for greater characterization of how NOs in surface waters affects N cycling
and has led to improved understanding of how factors such as seasonality and
contribution of snowmelt affect NO3 concentrations.
7.1.2.2. Sulfate in Surface Waters
Measurements of SO42 concentration in surface waters provide important information
regarding the probable extent of cation leaching in soils and how SO42 concentrations
relate to ambient and past levels of atmospheric S deposition. The 2008 ISA documented
widespread declines in surface water SO42 concentration in response to decreasing S
deposition over previous decades. Assessments of acidifying deposition effects reported
in the 2008 ISA dating from the 1980s showed SO42 to be the primary strong acid anion
in most, but not all, acid-sensitive waters in the U.S. (Webb et al.. 2004; Driscoll et al..
2001b; Driscoll et al.. 1988; Driscoll and Newton. 1985). At the time of the 2008 ISA,
available data indicated that before the peak of S emissions in the 1970s and 1980s, SO42
concentrations in surface waters increased in response to S deposition. After the
emissions peak, there were decreasing regional trends in SO42 surface water
concentrations in the 1980s, 1990s, and thereafter, particularly in the Northeast. In some
regions, especially the Blue Ridge Mountains region in Virginia, surface water SO42
remained relatively stable even as emissions declined, due to changes in S adsorption on
soils and the release of historically deposited S from soils into surface water. The rate of
surface water response was variable by watershed, and some model results suggested that
chemical recovery may be delayed as adsorption decreases and accumulated S leaches
from watersheds, even as emissions and deposition decline. Recent studies on SO42
concentration in surface waters continue to show evidence of chemical recovery
consistent with reductions in S deposition (Appendix 7.1.5.1').
Drought has been shown in some situations, where geological S sources are prevalent, to
be a partial cause of lake acidification. This was documented at lakes in the
Murry-Darling Basin in Australia by Li et al. (2017b). Pyritic sediment rewetting after
the drought (2007-2010) caused an extreme decrease in the water pH around the lake
margins from above 7 to near 3.
June 2018
7-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
7.1.2.3. Nitrogen Transformations
Nitrogen in surface waters can undergo various chemical transformations in the water
column (Appendix 7.1.2.3.1) and sediments (Appendix 7.1.2.3.2'). It was well known at
the time of the 2008 ISA that nitrification and denitrification are quantitatively important
portions of the N cycle and that these processes can be influenced by atmospheric inputs
of oxidized and reduced N. Nitrification can be important even when N deposition is low
(cf.. Fenn et al.. 2005). More recent research has further substantiated earlier findings and
provided additional quantitative context. At some locations, new research suggests that
denitrification may play a larger role than was previously recognized in removing
oxidized N from the watershed.
7.1.2.3.1.	Water Column Transformations
Much of the deposited NH4+ is either taken up by vegetation or nitrified to NOs . which is
more mobile in soils than NH44". The NO3 . in turn, can be leached to drainage water or
denitrified and released back to the atmosphere as gaseous N2O or N2. During the
nitrification process, NH44" is oxidized to NO3 and the NO3 that is produced can
contribute to acidification of soil and drainage waters. The 2008 ISA concluded that
nitrification in soil solution is stimulated at soil C:N ratios less than about 20 to 25. (U.S.
EPA. 2008a; Aber etal.. 2003). Since the 2008 ISA, additional isotope, modeling,
observational, and experiment studies have further characterized denitrification
processes.
Only a small portion of the N added to the land surface by human activities is carried by
stream flow to estuaries and the ocean (Bovcr et al.. 2002; Howarth et al.. 1996a).
Streams have been shown to provide an important ecosystem service by acting as N sinks
(Mulholland et al.. 2008). Denitrification is a critical process that removes N from stream
water. Changes in N2 concentration over space or time have been used to estimate the
denitrification rate at the scale of stream reaches (Mccutchan et al.. 2003; Laursen and
Seitzinger. 2002) to large river basins (Mulholland et al.. 2008; Alexander et al.. 2000).
Nitrous oxide (N2O), which is emitted to the atmosphere during the process of
denitrification, is a potent greenhouse gas. Beaulieu et al. (2011) presented results of 15N
tracer addition to 72 headwater streams draining multiple land uses across the U.S.
Denitrification in the streams produced N2O at rates that increased with stream NO3
concentration. The study streams were mostly sources of N2O to the atmosphere. Results
of this study reiterated previous findings that suggested that the process of denitrification
may be quantitatively important. In another isotopic study, Mulholland et al. (2009)
measured denitrification rates using 15N tracer addition to 49 streams, including
June 2018
7-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
reference, agricultural-impacted, and suburban/urban streams. The fraction of total NOs
removed from stream water by denitrification ranged from 0.5 to 100%, with a median of
16%. Removal was related to NH44" concentration and the ecosystem respiration rate.
Although the areal denitrification rate increased with increasing NO3 concentration, the
efficiency of NO3 removal from water via denitrification declined. This resulted in a
smaller proportion of stream water NO3 load removed over a given length of stream at
higher N loading.
Freshwater nutrient cycling studies have also focused some attention on N fixation. In
Waco Reservoir, TX, Scott et al. (2008) measured N fixation over a 19-month period and
linked those data with nutrient-loading estimates derived from a physically based
watershed model. Readily available topographic, soil, land cover, effluent discharge, and
climate data were used in the Soil and Water Assessment Tool (SWAT) model in order to
derive estimates of nutrient loading from the watershed to the reservoir. Results
suggested that human activities in the watershed can exert significant control over
planktonic N fixation.
Sediment Nitrogen Transformations
The deposition and transport of N derived from human sources is partially ameliorated by
bacterially mediated denitrification in sediments. The conversion of NO3 to N2 gas
permanently removes N from the watershed (Scitzingcr et al.. 2006; Galloway et al..
2003). It has been estimated that denitrification in lakes may remove up to about 30% of
the inputs to surface waters (McCrackin and Elser. 2012; Harrison et al.. 2009; Wollheim
et al.. 2008).
Residence time of water in the sediments and hydraulics can be important controls on
nutrient uptake in headwater streams. Drummond et al. (2016) characterized water
transient storage zones and their effects on nutrient uptake in the sediments of two
headwater streams in Spain. These zones represented regions of slow-moving water and
temporary water retention in stream sediments. Modeled exchange between the water
column and retention zones explained more than 40% of the variation in NH4+ uptake.
A study by Bellinger etal. (2014) provided additional evidence that bacterially mediated
denitrification in lake sediments can partly ameliorate the effects of N loading from the
atmosphere by permanently removing some of the N inputs. McCrackin and Elser (2012)
also measured denitrification in sediments, in this case from lakes in the Colorado Rocky
Mountains. The study lakes received either elevated (357-571 eq N/ha/yr) or low
(<142 eq N/ha/yr) inputs of atmospheric N deposition. The NO3 -N concentration was
significantly higher in high-deposition lakes (0.1582 mg/L) compared to low deposition
7-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
lakes (0.0462 mg/L). The researchers concluded that the sampled lakes were capable of
removing a significant portion of N inputs via denitrification in the lake sediment. They
found no difference between high- and low-deposition lakes in the extent to which
chronic N loading has altered sediment denitrification capacity. The abundance of
denitrifying bacteria in this study was largely related to light availability. An earlier study
by McCrackin and Elser (2010) measured rates of denitrification and N2O production
during denitrification in the sediments of 32 Norwegian lakes at the high and the low
ends of a gradient of atmospheric N deposition. They also found that denitrification
varied with N loading. The findings also suggested that lake sediments may be an
important source of N2O emissions, which play a role in climate warming, particularly in
areas that are subject to elevated N deposition levels. Results of new research by
McCrackin and Elser (2012) support growing evidence that lake sediments can play
important roles in N removal, although it appears that recent levels of N deposition have
not altered the abundance of denitrifying bacteria or saturated the capacity for sediment
denitrification in Rocky Mountain lakes.
Organic matter can also influence the rate of denitrification in sediments. Fork and
Heffernan (2014) investigated how DOC derived from terrestrial ecosystems affects the
rate of denitrification in river sediments. They examined the extent to which higher
concentrations of DOC in black water rivers stimulate or inhibit denitrification. Results
supported the hypothesis that terrestrially derived DOC might indirectly inhibit
denitrification via a decrease in autochthonous production. Therefore, changes in DOC
concentration might change the ability of inland waterways to remove reactive N from
the aquatic ecosystem.
7.1.2.4. Sulfur Transformations
Studies of S cycling reported in the 2008 ISA emphasized the importance of S adsorption
and desorption and their interactions with soil pH (Appendix 4). The importance of S
adsorption on soils in the southern Appalachian Mountains was further confirmed by S
budget studies reported by Rice et al. (2014). In addition, internal watershed sources of
S042 (e.g., S mineralized from soil organic matter), which were previously relatively
minor sources to surface water in the northeastern U.S., have become proportionately
more important as S deposition has declined (Appendix 4).
Both chronic and episodic leaching of SO42 from terrestrial ecosystems influence surface
water acidification. The literature reviewed in the 2008 ISA did not fully address how
terrestrial S cycling affects water sulfate concentrations. More recently, Rice et al. (2014)
calculated SO42 mass balances for 27 watersheds in the Appalachian Mountain region.
June 2018
7-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Results suggested that many watersheds where SO42 inputs are currently retained will
begin releasing more SO42 to drainage water in the near future. Dewalle et al. (2016)
evaluated Appalachian forested watersheds and estimated the lag times between changes
S deposition and consequent watershed responses. Lag times were generally consistent
and significant for S in the forest basins that demonstrated relatively high soil S retention
and high N retention. Strock et al. (2016) investigated the effects of extreme (wet or dry)
weather on the chemistry of 84 lakes across the northeastern U.S. The average
differences in lake water SO42 concentrations were about 2-3 j^icq/L higher in 2001 (dry
year) as compared with the average observed during the entire study period 1990-2010.
This has implications for S cycling as the climate changes.
Some of the SO42 leached from soil, as well as the SOx deposited directly into surface
water, is reduced and retained in aquatic sediments as hydrogen sulfide, especially in
wetlands. However, S stored in reduced form in sediments can be subsequently
reoxidized and become available for down-gradient transport as SO42 during periods of
high discharge, particularly during hydrologic events that follow periods of drought. The
leaching of SO42 then can contribute to a variety of ecological effects (Appendix 8 and
Appendix 12). When SO42 is released from catchment soils to drainage water, it is
accompanied by an equivalent amount of cationic counter-charge in the form of acidic
(H+, Al3+) or basic (Ca2+, Mg2+, K+, and Na+) cations (U.S. EPA. 2008a). affecting the
concentrations of parameters addressed in the following sections of this report.
7.1.2.5. Surface Water pH
Surface water pH is commonly used as an indicator of acidification. This was the case in
the 2008 ISA, and nothing has changed in that regard in more recent years. It correlates
with other biologically important components of surface water acid-base chemistry,
including estimates of organic acidity and concentrations of inorganic Al and Ca2+. Low
pH can have direct toxic effects on aquatic species IYU.S. EPA. 2008a; Driscoll et al..
2001b); Appendix 8.31 and this was widely understood at the time of the 2008 ISA. Low
pH can disturb normal ion osmoregulation in aquatic biota. Threshold pH levels for
adverse biological effects are described in Appendix 8 and Table 8-2. A pH value of 6.0
is often considered the level below which biota are at increased risk from acidification,
but some waters can have pH below this threshold in the absence of acidic deposition.
This is most commonly caused by relatively high levels of natural organic acidity. Below
pH 5.5, inorganic Al often becomes the greatest threat to aquatic biota, especially in
low-DOC waters. In the 2008 ISA, increasing trends in pH in surface waters in the
northeastern U.S. were common through the 1990s up to 2004 and have continued in
June 2018
7-23
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
more recent times at many locations (Appendix 7.1.5.1). Rates of change have generally
been relatively small.
Surface water pH (as well as other parameters) is sensitive to hydrological conditions at
the time of sampling. Burns et al. (2008b) sampled 12 Catskill mountain streams within
the Neversink River watershed in New York for stream chemistry. The measured pH
values in 2003 were generally lower than those in 1987 and this was attributed to higher
stream flow during the summer of 2003, rather than to chronic acidification during the
intervening period. There were no significant differences in biota observed between 1987
and 2003. Although surface water pH is a common alternative to ANC as an indicator of
acidification, at pH values above about 6.0, pH is of less value as an indicator of either
sensitivity to acidification or level of effect. In addition, pH measurements (especially at
these higher values) are sensitive to levels of dissolved CO2 in the water, which is
influenced by plant root respiration.
7.1.2.6. Surface Water Acid Neutralizing Capacity
The most widely used measure of surface water acidification, and subsequent recovery
under reduced acid deposition, is ANC. Most aquatic critical load (CL) studies conducted
in the U.S. have used surface water ANC as the principal metric of water quality change
in response to changes in acidic deposition (Appendix 7.1.5.2). It is typically either
determined by Gran titration in a laboratory (titrated ANC) or calculated from the charge
balance (calculated ANC or CALK). Titrated ANC is useful because it reflects the ANC
of the complete chemical system, which is typically decreased by acidic deposition in
acid-sensitive landscapes. The ANC is associated with the surface water constituents that
directly cause or reduce acidity-related stress, in particular pH, Ca2+, and inorganic Al
concentrations. The ANC is generally a more stable measurement than pH because ANC
is insensitive to changes in CO2 and it reflects sensitivity and effects of acidification in a
linear fashion across the full range of ANC values. Therefore, ANC is usually the
preferred indicator for assessment of surface water acidification and recovery. Both
titrated and calculated ANC values are commonly determined in studies aimed at
resource characterization or long-term monitoring. Models simulate calculated ANC. The
two measures can differ greatly, depending mainly on the amount of organic acidity and
dissolved Al in the water. The BCS (Lawrence et al.. 2007) is an alternate index to ANC
that integrates acid-base status and accounts for the influence of natural organic acidity.
As described in the 2008 ISA, ANC is typically used as the primary chemical indicator
for assessing past effects of acidifying deposition on aquatic biota and the recovery
expected from decreasing acidic atmospheric deposition (Sullivan et al.. 2006a; Aber et
June 2018
7-24
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
al.. 2003; Bulger et al.. 2000). Notably, the ANC level that reflects recovery of pH or
inorganic Al may differ between the acidification and recovery phases (Hcsthagcn et al..
2008)(Appendix 7.1.5.1). In general, ANC measures over the past two to three decades in
the northeastern U.S. (cf.. Driscoll et al.. 2016) suggest rather modest increases in ANC
in many surface waters in response to large decreases in acidic deposition and surface
water SO42 concentration (Appendix 7.1.5.1).
Information on biological indicators of acidification such as fish species richness that are
associated with changes in ANC are presented in Appendix 8. There is often a positive
relationship between ANC and number of fish species, at least for ANC values between
about 0 and 50 to 100 |icq/L (Cosby et al.. 2006; Sullivan et al.. 2006a; Bulger et al..
1999). Lakes and streams having ANC <0 j^icq/L generally do not support fish. Loss of
fish species seems to occur with decreases in ANC below a threshold of approximately
50 to 100 (ieq/L (Sullivan et al.. 2006a).
Sensitive water bodies can be defined as those that have ANC of 100 |icq/L or less.
Sensitivity increases with further decreases in ANC. Al mobilization is largely confined,
however, to waters that have pH less than about 5.5, which corresponds with ANC in the
range of about 10 to 30 j^ieq/L in low-DOC to moderate-DOC (less than about 400 |iM)
waters in the Northeast. Therefore, inorganic Al is not a useful indicator of acidification
or chemical recovery in waters that have ANC higher than about 10-30 j^icq/L. Thus,
evaluation of improvement in biologically relevant water chemistry in response to
decreases in acidic deposition should perhaps include assessment of response using both
the ANC and inorganic Al metrics.
Povak et al. (2013) reported the ANC of over 900 streams in the southern Appalachian
Mountains, and estimated ANC at other stream segments throughout the region using
observed relationships between ANC and watershed characteristics. Low stream ANC
was commonly found at locations that exhibited siliciclastic geology; cool, short, and
moist growing seasons; high clay soil content; low soil pH; and small forested
watersheds.
Sullivan (2017) constructed a map of surface water ANC across the U.S. that included
nearly 20,000 unique locations sampled between 1980 and 2011 (Figure 8-11). Samples
expected to be strongly influenced by acid mine drainage, sea salt spray, or road salt
application were excluded. He found 6,065 sites that had ANC <100 j^ieq/L. Waters
having ANC <0 j^ieq/L were mostly restricted to northern New York, New England, the
Appalachian Mountain chain, upper Midwest, and Florida. In addition, low, but positive,
ANC values were found in high-elevation portions of the West and parts of Arkansas and
the Gulf states. These spatial patterns in surface water ANC are thought to mainly reflect
the influence of soil base cation supply on ANC.
7-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
7.1.2.7. Base Cation Concentrations
Quantitatively, the most important component of the overall surface water acidification
and chemical recovery response to atmospherically deposited S and N (and associated
S042 and N03 concentrations) has been a change in base cation supply (Charles and
Christie. 1991). As stated in the 2008 ISA, decreases in base cation concentrations in
surface waters in the eastern U.S. over the past two to three decades have been ubiquitous
and closely tied to trends in SO42 concentrations in surface waters. In most regions, rates
of decrease for base cations have been similar to those for SO42 plus NO3 . with the
exception of streams in western Virginia and in Shenandoah National Park, which are
strongly affected by SO42 adsorption on soils. The 2008 ISA concluded that acidifying
deposition has been an important cause of decreasing amounts of exchangeable base
cations on soils. This influences base cation leaching to surface waters and soil buffering
of deposition acidity. Surface water base cation concentrations increase with acidification
and generally decrease under reduced levels of acidic deposition. However, in some
watersheds that have acidified, concentrations of Ca2+ and other base cations have
become substantially reduced from likely preindustrial levels, in response to depletion by
many years of acidic deposition. This base cation depletion constrains surface water ANC
and pH recovery, as described in the 2008 ISA. Recent results have further corroborated
these earlier findings. More recent studies have included experiments (liming), modeling,
and gradient studies with geographical focus on the Adirondack Mountains, southern
Appalachian Mountains, Rocky Mountains, and the Hubbard Brook Experimental Forest
(HBEF) in New Hampshire.
Base cations are contributed to soils and surface waters through weathering and
atmospheric base cation deposition (Appendix 4). Estimates of base cation weathering
(BCw)largely control simulated base cation concentrations in drainage water and are
needed for evaluating CLs of surface water acidity using steady-state models
(Appendix 8). McDonnell et al. (2012) developed an approach for regionalizing BC„
using regionally specific empirical relationships. The dynamic model MAGIC was used
to calibrate BC„ in 92 watersheds distributed across the southern Appalachian
Mountains. About one-third of the study region had BC„ estimates that were less than
1,000 eq/ha/year, with lowest values for watersheds located in national parks and
wilderness areas.
Base cation surplus is defined as the difference between the summed concentrations of
base cations (Ca, Mg, Na, K) and strongly acidic inorganic anions (SO42 , NO3 .
chloride), minus an estimate of the strongly acidic organic anions estimated from
dissolved organic C and an assumed charge density. These strongly acidic organic anions
are dissociated at low pH, and function essentially as mineral acid anions in terms of their
June 2018
7-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
effect on ANC. The BCS is an alternative statistic to ANC for use mainly in waters that
have appreciable natural organic acidity.
In high-elevation portions of three watersheds in the southern Appalachian Mountains,
Knoepp et al. (2016) found that stream acid-base chemistry was related to the
concentrations of N and Al in the soil O-horizon and to the total amount of C and Ca in
the soil. Each watershed contained four to six first-order catchments with moderate to
high levels of acid sensitivity (ANC 11-50 (ieq/L). Differences across catchments in
stream ANC, pH, and the ratio of Ca:Al concentrations were also significantly correlated
with watershed vegetation, as reflected in basal area, tree height, and diameter at breast
height.
In the Muskoka River Watershed in Ontario, Canada, further improvement in lake ANC
and pH might be limited by Ca depletion, which can be made worse by tree harvesting
(Rcid and Watmough. 2016). More than 60% of the tree Ca is commonly found in the
bark and boles, which are largely removed from the watershed during timber extraction.
The authors estimated that timber harvesting at planned levels will cause approximately a
30% increase in the number of sampled lakes that decline to Ca levels below 1 mg/L.
Changes in acidic deposition exert complex effects on the concentration of Ca in surface
waters. During the early phases of soil and water acidification, acidic deposition increases
the leaching of Ca from soil to drainage water. This Ca leaching causes increased
concentrations of Ca in streams and lakes. In base-poor watersheds having thin soils, the
rate of Ca leaching can exceed the resupply via weathering and atmospheric input. Over
time, soil base saturation decreases, and the concentration of Ca in runoff decreases,
perhaps to levels lower than what existed prior to the advent of acidic deposition.
Decreases in acidic deposition will further accentuate this Ca depletion. Decline in lake
water Ca concentration has severe consequences for some species of zooplankton,
especially Daphnia species that have high Ca requirements. Jeziorski et al. (2008)
documented major reductions in abundance of Ca rich Daphnia spp., keystone herbivores
in pelagic food webs, in association with decreases in the concentration of Ca in lake
water. They reported that a high proportion of Canadian Shield lakes had Ca
concentration near or below the threshold level (1.5 mg/L) for decreased survival and
fecundity in laboratory studies. Ecological impacts of these changes on food webs may be
substantial.
Increased atmospheric transport of dust enriched with nutrients, metals, and base cations
to alpine watersheds in the western U.S. has been shown to alter lake biogeochemical
processes. Ballantvne et al. (2011) examined temporal trends in dust deposition from
sediment cores collected from two lakes in Colorado with distinct physical and chemical
properties. Recent increases in dust deposition and its enrichment in various elements,
7-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
including Ca, has contributed base cations to drainage water and altered the
biogeochemistry of the two study lakes.
Liming studies confirm ecosystem responses when depleted base cation pools in soils are
restored by addition of base cations experimentally or for management purposes. In a
base cation addition study (45,000 kg of calcium silicate [wollastonite, CaSiOs |) in
HBEF, increases in the concentrations of Ca and decreases in the concentrations of H+
and inorganic A1 in stream water were observed and steam ANC increased (C'ho ct al..
2012). By the end of 2010, an estimated 3 to 5% of added Ca was exported in stream
water and the Ca retained in the watershed was detected in lower soil horizons or taken
up by vegetation (Shao et al.. 2016; Johnson et al.. 2014). Exchangeable Ca increased
significantly and exchangeable Al decreased significantly in the organic and upper
mineral soils over the years following wollastonite addition (Johnson et al.. 2014).
7.1.2.8. Surface Water Aluminum
As stated in the 2008 ISA, the concentration of dissolved inorganic monomeric Al in
surface waters is an especially useful indicator of acidifying deposition effects because
(1) it is toxic to many species of aquatic biota (Appendix 8) and (2) it generally does not
leach from soils to surface waters in the absence of input of strong acids (e.g., sulfuric
acid, nitric acid, strong organic acid anions)(Lawrence et al.. 2007; Driscoll et al.. 1988).
It has well-documented effects on aquatic biota at specific thresholds (Appendix 8). In
the 2008 ISA, limited data suggested that some acid-sensitive surface waters in regions of
the northeastern U.S. have elevated inorganic Al concentrations, which have been
induced by years of acidifying deposition and pose threats to aquatic life. Since the 2008
ISA, several monitoring studies have reported decreases in inorganic Al suggestive of
chemical recovery of surface waters lYBaldigo et al.. 2016; Driscoll et al.. 2016; Strock et
al.. 2014; Warbv et al.. 2008) (Appendix 7.1.5.1)1.
7.1.2.9. Surface Water Dissolved Organic Carbon Concentration
An especially important, and relatively new, area of research focus in the context of
recovery from surface water acidification has been the observed increase in DOC or total
organic carbon (TOC) concentration in many surface waters recovering from
acidification. This has implications with respect to water toxicity and cycling of C, Al,
and N. Changes in lake DOC are important because DOC helps to regulate biological,
chemical, and physical lake characteristics by adding C, changing pH, affecting nutrient
cycling, and changing the availability of toxic metals, including Al. Lake DOC regulates
June 2018
7-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
UV light penetration into the water column and the amount and depth of
photosynthetically active radiation (PAR), which affects the lake mixing depth, especially
in small lakes. Increased DOC contributes to shallower mixing depth and greater
fluctuation in diel temperature and may affect lake nutrient status (Gerson et al.. 2016).
Increases over the past two to three decades in lake DOC in the northeastern U.S. have
likely been part of the chemical recovery from previous lake acidification (SanClements
et al.. 2012; Monteith et al.. 2007) and/or caused or exacerbated by changes in climate,
including precipitation patterns (Couture et al.. 2012; Wevhenmever and Karlsson. 2009).
The 2008 ISA reported widely observed increased concentrations of DOC or TOC in
surface waters across North America and Europe and that these increases were at least
partly related to changes in atmospheric deposition of S and N. Thus, it has been
recognized for well over a decade that surface water organic carbon concentrations have
decreased to some extent in many water bodies that previously experienced water
acidification. Therefore, DOC concentration would likely increase with recovery. More
recent research on this topic has been diverse and has included experiments, observation,
isotope studies, and synthesis and integration work. Overall, they illustrate relatively
large increases in DOC with acidification recovery in some aquatic systems. This
response constrains the level of ANC, and especially pH, recovery, but decreases the
toxicity of dissolved Al by converting some of it from inorganic to organic forms
(Lawrence et al.. 2013). As a result of this widely observed pattern, surface water DOC
concentration has become an important focus of water acidification and recovery research
(Appendix 7.1.5.1).
DOC is comprised of a diverse mix of organic matter and functional groups. Wood et al.
(2011) noted that dark, aromatic-rich organic matter with relatively high humic content
and of allochthonous origin may be more effective in ameliorating metal bioavailability
and toxicity than other organic materials. In addition, part of the protection provided by
DOC may involve physiological mechanisms other than metal complexation (Wood et
al.. 2011). Soil mechanisms that contribute to higher DOC in surface waters are discussed
in Appendix 4. At the Niwot Ridge Long-Term Ecological Research site in Colorado,
Miller and Mcknight (2015) reported downstream transport of dissolved organic matter
(DOM) that had been produced in alpine lakes during low-flow periods.
Recent research suggested that increases in lake DOC may have changed key functions of
affected lakes, causing decreased vertical distribution of phytoplankton and changes in
phytoplankton species composition. This was investigated by Brown et al. (2017). who
demonstrated that increases in lake DOC can potentially change the thermal structure of
lakes, but that effects are variable. They analyzed fossil diatom remains in sediment cores
collected from three pairs of small remote Maine lakes. Each pair consisted of one lake
7-29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
that showed recent increases in DOC since the early 1990s and a similar lake that did not
show increasing DOC. The focus was on changes in the relative abundances of diatom
taxa that were known to reflect effects on thermal stratification, specifically Discostella
stelligera and Aulacoseira spp. Saros et al. (2012) had constructed a diatom inference
model for reconstructing past lake mixing depth based on the relative abundance of D.
stelligera, which prefers shallower mixing depth. Stone etal. (2016) expanded this model
and included species of Aulacoseira. Brown et al. (2017) hypothesized that those lakes
that had experienced recent DOC increases (shown in Figure 7-3) would show increased
diatom turnover and increases in the relative abundance of D. stelligera and Aulacoseira
spp. since the 1990s. The three study lakes that had no increase in DOC also showed
minimal change in the identified sensitive taxa. The three lakes that experienced recent
increases in DOC showed variable diatom responses. Results were interpreted as being
indicative of the potential of DOC to change the physical and biological structure of lakes
that have experienced past or ongoing chemical recovery from prior lake acidification.
The largest changes in diatom taxa occurred in two of the lakes that had experienced
recent increases in DOC.
DOC effects on N cycling have been further elucidated since the 2008 ISA. Rodriguez-
Cardona et al. (2016) and Wvmore et al. (2016) suggested greater NOs removal from
drainage water under conditions of relatively high DOC and high ratio of DOC: NO,
concentration. Results of the study of Fork and Heffernan (2014) further suggested that
the indirect effects of DOC on light will be important in determining how N cycling
responds to increasing concentrations of DOC in surface waters. Under conditions of
increased DOC, the quality of DOM may change, as may its binding with dissolved Al
(Fakhraei and Driscoll. 2015). Dynamics of DOC may also have consequences for
eutrophication of downstream ecosystems, but interactions appear to be complex.
7-30
DRAFT: Do Not Cite or Quote

-------
(A) Bracey*
(B) Salmon
O)
E,
o
o
o
1990 1995 2000 2005 2010 2015 1990 1995 2000 2005 2010 2015
(C) Jellison*
(D) Second
C5>
Ł
O
o
~
1990 1995 2000 2005 2010 2015 1990 1995 2000 2005 2010 2015
(E) Little Long*
(F) Tilden
oi
E.
O
O
Q
1990 1995 2000 2005 2010 2015 1990 1995 2000 2005 2010 2015
Year	Year
Notes: linear regression lines are shown for significant trends (p <0.05). Asterisk indicates lakes with a significant DOC increase.
Source: Brown et al. (2017).
Figure 7-3 Dissolved organic carbon (DOC) concentrations from 1993 to
2013 for Bracey Pond (a), Salmon Pond (b), Jellison Pond (c),
Second Pond (d), Little Long Pond (e), and Tilden Pond (f) in
Maine.
1	Short-term studies to characterize acid-base chemistry have provided information on
2	surface water quality and biogeochemistry of organic materials. Kang and Mitchell
3	(2013) studied spatial and temporal variation in the quantity and quality of DOC, N, and
4	S in Adirondack Mountain surface waters over a 14-month period. This investigation
June 2018	7-31	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
included the sampling of two upland streams, two wetland-influenced streams, and one
lake outlet (Arbutus Lake). The DOC and dissolved organic N (DON) concentrations
increased as water was transported through wetland areas. Results highlighted the value
of applying multiple approaches for understanding the biogeochemistry of DOM. The
acid-base chemistry of DOC has been further investigated in several other studies. Porcal
et al. (2009) confirmed that DOC concentrations in lakes and streams throughout much of
Europe and North America have increased over the last three decades. Possible reasons
for this change proposed by Porcal et al. (2009) included increased atmospheric CO2
concentration, climate warming, decreased S deposition (and associated changes in water
pH and ionic strength), and hydrologic changes caused by drought and precipitation. Any
changes in DOC concentration or properties will impact the acid-base chemistry of
surface waters and perhaps the composition of aquatic biota. In a laboratory study, Al-
Reasi et al. (2013) performed titrations for a range of freshwater DOM isolates. In
general, the proton site density (Lt) exhibited maxima near proton binding constants
(pKa) values of 3.5 and 10, reflecting the presence of both strong and weak organic acids.
The Proton Binding Index parameter was described to summarize the chemical reactivity
of DOM based on the pKa values and the Lt. Results of this study provided a rationale for
describing the protection provided by DOM against metal, including inorganic Al,
toxicity. Recent additional research has also focused on improving scientific
understanding of variation in the importance of organic versus mineral acids in
controlling the acidity of soil water and surface water. Chapman et al. (2008) analyzed
seasonality in soil water and surface water chemistry in two U.K. headwater watersheds.
Strong seasonal patterns were observed for both ANC and DOC. Acidic deposition
controlled acidity during winter, whereas organic acids were more important during
summer. The observed increase in DOC concentration in drainage water over the
previous two decades was attributed mainly to climate warming, increased CO2, and/or
decreased acidic deposition.
SanClements et al. (2012) used a chemical signature of terrestrial DOM to investigate the
extent to which increased DOC concentrations in surface water since about 1993 in the
northeastern U.S. have been driven by decreasing acidic deposition and increasing
solubility of soil organic matter. They used fluorescence spectroscopy to characterize the
quality of DOM in stored samples that had been collected from nine acid-sensitive lakes
in Maine. Decreases in lake water SO42 concentration were associated with increases in
DOC concentration and a shift during ecosystem recovery from microbial to terrestrially
derived DOM. Changes in the quality or quantity of DOM can affect aquatic ecosystem
functions. Dissolved organic matter is important for providing food for microbes,
attenuating light, buffering pH, binding Al, and controlling the cycling of nutrients
(SanClements et al.. 2012).
7-32
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
The amount and composition of DOC mobilized from upland soils to surface waters is an
important link in the global C cycle. The relationship between ultraviolet (UV)
absorbance and DOC concentration has been shown to reflect the aromaticity of the
DOC. Dawson et al. (2009) reported a change in correspondence between UV absorbance
and DOC concentration over a period of 22 years at two moorland watersheds in
Scotland. The DOC concentration has increased over time, while the proportion of
hydrophobic DOC has decreased. A model analysis by Dawson et al. (2009) suggested
that S deposition was the only factor that could reasonably explain the observed
long-term DOC trend at both sites.
Valinia et al. (2015) used visible near-infrared spectroscopy of lake sediments to
reconstruct the reference condition TOC in Swedish lakes. Long-term monitoring data
were used to simulate recent changes in TOC and then analyzed using two empirical
models. The first predicted historical TOC trends between reference conditions in the
year 1860 and peak acidification in approximately the year 1980. The second model
predicted TOC between 1988 and 2012 in conjunction with partial chemical recovery
from acidification. The models were driven by lake and watershed area, the amount of
wetlands, historical S deposition, and current water chemistry. The researchers estimated
that the present-day TOC concentrations are similar to reconstructed reference conditions
in Swedish lakes. Thus, it was deemed unlikely that the TOC concentrations in Swedish
lakes will continue to increase with continued controls on acidic deposition.
In a monitoring study from the Czech Republic, Hruska et al. (2009) reported high DOC
in two watersheds: 18.8 mg/L and 20.2 mg/L at the acidic and base-rich watersheds,
respectively. Between 1993 and 2007, the concentrations of DOC in stream water
increased by about 65% at both streams. Increases in stream water DOC were associated
with small increases in stream pH, but large decreases in ionic strength due to declining
acidic deposition. Although neither of the catchments showed changes in soil water pH,
DOC concentrations in soil water tripled. Hruska et al. (2009) concluded that the change
in ionic strength of soil water and stream water, rather than acidity, was the major cause
of increased stream water DOC. Changes in temperature, precipitation, and discharge
during the study suggested that climate change may be an additional important control of
DOC dynamics. The role of ionic strength in modifying temporal patterns of surface
water DOC in the U.S. has not been studied at the time of this writing. Evans et al. (2017)
also documented decreased lake DOC in response to reacidification of a moorland lake
catchment in the U.K. In addition, the quality of the DOC changed towards higher levels
of aromatic organic compounds and an increase in particulate organic C.
7-33
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
7.1.2.10.
Climate Modification of Ecosystem Response to Nitrogen and Sulfur
Surface water chemistry and chemical recovery (Appendix 7.1.5.1) of fresh waters are
occurring within the context of projected changes in annual mean temperature and
magnitude of precipitation associated with climate change (G reaver et al.. 2016). In acid
sensitive regions, altered hydrologic regimes are likely to affect weathering rate of base
cations, lake water levels, and organic matter inputs to catchments (Adrian et al.. 2009;
Porcal et al.. 2009). Projected shifts in runoff, timing, and quantity of flushing will alter
the frequency and duration of episodic events and the concentrations of nutrients and
chemical indicators in surface waters (Adrian et al.. 2009; Whitehead et al.. 2009).
Extreme weather years (wet or dry) can shift water chemistry responses such as DOC and
S042 concentrations (Strock et al.. 2016). Increased loading of DOC to surface waters,
attributed in part to increasing temperatures and changes in precipitation, may affect
biogeochemical processes and nutrient availability (Daggett et al.. 2015; Zhang etal..
2010; Wevhenmever and Karlsson. 2009). Appendix 13 includes a more detailed
discussion of how climate (e.g., temperature and precipitation) modifies ecosystem
response to acidification.
7.1.3. Freshwater Monitoring and Databases
Long-term monitoring of surface water chemistry over time has enabled a greater
understanding of ecosystem response to deposition of N and S, including chemical and
biological responses. Monitoring data inform determination and quantification of
temporal trends and many monitoring studies for acidification have been ongoing for one
or two decades, in some cases longer. Such data were available and incorporated into the
2008 ISA. Since that time, nearly a decade's worth of additional data have been added to
some of these databases. This is noteworthy because short-term temporal variability can
mask small changes that are part of a long-term trend. The availability of these additional
data facilitates trend detection now, compared with 2008. A number of monitoring
studies have been conducted or continued in recent years that document nutrient
dynamics in water bodies in response to N inputs. Recent evidence that P deposition is
increasing (Appendix 9.1.1.4) may contribute to total nutrient loading and affect shifts in
lake nutrient status.
7.1.3.1. Acidification
A number of freshwater monitoring studies have been conducted using data collected
over the previous one to three decades that have documented ecosystem chemical damage
June 2018
7-34
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
and recovery (Appendix 7.1.5.1) caused by acidifying deposition of S and/or N. Many of
these studies have been conducted in the U.S., especially in the Northeast and the
southern Appalachian Mountains (Table 7-3). including a number of studies since 2008.
Two surface water chemistry monitoring programs that have been especially useful to
inform the assessment of aquatic ecosystem responses to changes in acidic atmospheric
deposition have been Temporally Integrated Monitoring of Ecosystems (TIME)
(Stoddard et al.. 1996) which is no longer operating, and the U.S. EPA Long-Term
Monitoring (LTM) project (Stoddard et al.. 1998; Ford et al.. 1993). These U.S. EPA
monitoring efforts focus on portions of the U.S. most affected by the acidifying influence
of S and N deposition, including lakes in the Adirondack Mountains of New York and in
New England, and streams in the Northern Appalachian Plateau and the Blue Ridge
physiographic provinces in Virginia and West Virginia. Both projects are operated
cooperatively with numerous collaborators in state agencies, academic institutions, and
various federal agencies.
The Adirondack Lakes Survey Corporation (ALSC) has been monitoring Adirondack
lakes for about 30 years. This work has mainly focused on acid-base chemistry, but has
also involved some fish monitoring and measurement of parameters relevant to nutrient
enrichment. Monitoring data have shown some chemical recovery from lake
acidification, reflected in increased pH and ANC and decreased inorganic Al
concentrations (Appendix 7.1.5.1). HBEF in NH also has several decades of monitoring
data. The central and southern Appalachian Mountains region is important because (1) it
contains an abundance of low-ANC streams situated on base-poor geology,
(2) atmospheric S and N deposition have been high, (3) S adsorption on soils complicates
acidification/recovery responses, and (4) much of the acid-sensitive landscape is
managed as national park and wilderness area. New studies have been conducted of
stream acid-base chemistry throughout this region in West Virginia, Maryland, Virginia,
and Tennessee. More recent monitoring studies are highlighted below and in Table 7-3.
Since the early 2000s, U.S. EPA, together with states, tribes, other entities, and
individuals, have collaborated on a series of statistically representative surveys (National
Aquatic Resource Surveys [NARS]) of the nation's waters, including surveys of lakes
(U.S. EPA. 2016h. 2009b). streams (U.S. EPA. 2016i). wetlands (U.S. EPA. 2016i). and
coastal waters (U.S. EPA. 2016g). Based on standard sampling and analysis protocols
and consistent quality assurance, these surveys periodically assess the magnitude and
spatial extent of water quality issues and concerns across the U.S.
7-35
DRAFT: Do Not Cite or Quote

-------
Table 7-3 Monitoring and resurvey results of aquatic acidification and/or chemical recovery since the 2008
Integrated Science Assessment for Oxides of Nitrogen and Sulfur—Ecological Criteria.
Process
Acidification
Indicator
Nutrient
Enrichment
Indicator
Type of
Ecosystem
Region
Time
Period
Ambient
N/S
Deposition
kg/ha/yr
Effect of Deposition
Publication
Chemical
recovery
pH, ANC,
SO42", DOC
N/Ap
12 streams
Adirondack
Mts.
1980s-2008 Variable
On average, pH increased by
0.28; ANC increased 13 peq/L;
rate of decrease in stream
SO42" concentration was
2 pmol/L/yr between 1999 and
2008 at a high-DOC stream
and 0.73 pmol/L/yr at a
low-DOC stream.
Lawrence et al. (2011)
Chemical
recovery
pH, ANC,
S042", DOC,
NO3-, AN
NO3-
48 lakes
(16 lakes that
were part of
the original
ALTM
monitoring
project and
addition of
new lakes in
1992)
Adirondack
Mts.
1982-2015
All study lakes showed
significant decreases in the
concentration of SO42" in lake
water consistent with
reductions in S deposition.
Concentration of NO3"
declined at variable rates in 33
of the 48 study lakes.
Widespread increases in ANC
in 42 of the 48 study lakes and
in lab pH in 33 of the 48 study
lakes. AN decreased in 45 of
the 48 lakes. Dissolved
organic C increased in 29 of
the 48 lakes. Most of the
48 ALTM lakes showed
significant decreases in the
concentrations of base
cations. Eleven of the
16 original ALTM lakes
showed increases in lake ANC
between 1982 and 2013.
Driscoll et al. (2016)
June 2018
7-36
DRAFT: Do Not Cite or Quote

-------
Table 7-3 (Continued): Monitoring and resurvey results of aquatic acidification and/or chemical recovery since
the 2008 Integrated Science Assessment for Oxides of Nitrogen and Sulfur-Ecological
Criteria.
Process
Acidification
Indicator
Nutrient
Enrichment
Indicator
Type of
Ecosystem
Region
Time
Period
Ambient
N/S
Deposition
kg/ha/yr
Effect of Deposition
Publication
Chemical
recovery
ANC, DOC
NOs"
TIME lakes
Adirondack
Mts.
1991-2007 Variable
Percentage acidic lakes
decreased from 15.5 to 8.3%;
calculated ANC increased at
more than twice the rate as
Gran ANC, which was
attributed to increase in DOC.
Waller et al. (2012)
Chemical
recovery
S042", DOC N/Ap
Nine lakes Maine	1993-2009 Wet	Decreases in lake S042-
S = 6.2	correlated with increases in
(1980) to 1.5	DOC and a shift from microbial
(2010)	to terrestrially derived organic
N = 3	matter.
SanClements et al.
(2012)
Chemical
recovery
NO3-
NOs"
Lakes
Adirondack
Mts. and
New England
2000-2010 N/Av
Lake NO3" concentrations
declined at rate of
-0.05 [jeq/L/yr and there was
a shift to nontoxic (organic) Al.
Strock et al. (2014)
Chemical
recovery
Ali
N/Ap
Resurvey of
113 lakes
Northeastern
U.S.
1986-2001 N/Av
In 2001, only 7 lakes,
representing 130 lakes in
population, had Ali >2 pmol/L
(toxic threshold to brook trout),
compared with 20 sampled
lakes (representing 449 lakes
in population) in 1986.
Warbv et al. (2008)
June 2018
7-37
DRAFT: Do Not Cite or Quote

-------
Table 7-3 (Continued): Monitoring and resurvey results of aquatic acidification and/or chemical recovery since
the 2008 Integrated Science Assessment for Oxides of Nitrogen and Sulfur-Ecological
Criteria.
Process
Acidification
Indicator
Nutrient
Enrichment
Indicator
Type of
Ecosystem
Region
Time
Period
Ambient
N/S
Deposition
kg/ha/yr
Effect of Deposition
Publication
Chemical SCM2", NOs", NOs"
recovery ANC
Two streams Western MD 1990-2005 N/Av
Concentrations of SO42" in
stream water decreased at a
rate of about 3 peq/L/yr in
response to 34% reduction in
wet S deposition. Trends in
stream NO3" appeared to be
related to watershed factors,
especially forest disturbance.
Rate of ANC increase during
the period 1996-2005 was
about half the rate of the entire
study period, suggesting that
chemical recovery may be
slowing or coming to an end.
Eshleman etal. (2008)
Chemical
Ca2+, Mg2+,
NOs"
Two streams
Bear Brook,
1988-2006
Wet S = 6
Concentrations of Ca and Mg
Navratil etal. (2010)
recovery
SO42"

(manipulated
ME

(1987) to 2.3
in stream water decreased




and control)


(2006)
more than SO42"








concentration, causing stream








acidification in control stream.

Chemical
ANC
N/Ap
64 streams
Western
1987-2011
Wet S = 9
At most sites underlain by
Robison et al. (2013)
recovery



Virginia

(1980) to 3.7
base-poor bedrock, ANC







(2010)
decreased despite reductions







Wet N = 4.7
in S deposition. This response







(1980) to 3.0
was related to depletion of







(2010)
base cations.

S
S042"
NO3-
Stream
Noland
1991-2006
S = 28
Sulfur adsorption on soil is
Cai etal. (2010)
adsorption


watershed
Divide,


important on average. During





GSMNP, TN


large precipitation events,








SO42" was more mobile and








caused stream acidification.

June 2018
7-38
DRAFT: Do Not Cite or Quote

-------
Table 7-3 (Continued): Monitoring and resurvey results of aquatic acidification and/or chemical recovery since
the 2008 Integrated Science Assessment for Oxides of Nitrogen and Sulfur-Ecological
Criteria.
Process
Acidification
Indicator
Nutrient
Enrichment
Indicator
Type of
Ecosystem
Region
Time
Period
Ambient
N/S
Deposition
kg/ha/yr
Effect of Deposition
Publication
Pyrite
weathering
S042"
N/Ap
High
elevation
lakes
Colorado
1985-2008 N/Av
Lake SO42" concentration
decreased at a rate of-0.12 to
-0.27 [jeq/L/yr. Climate
warming appears to have
affected pyrite weathering and
lake SO42" concentration.
Mast etal. (2011)
ANC
production
Alkalinity, Ca2 N/Ap
Streams and Eastern U.S. Varies, at Varies	Acidifying deposition is one
rivers	least 25 yr	contributor along with
of	carbonate lithology and
monitoring	watershed topography to
data per site	significantly increasing
alkalinity trends (in 64% of
97 river and stream sites).
Most rapid rates of alkalization
occurred at sites with the
highest elevation and greatest
inputs of acidifying deposition.
Kaushal etal. (2013)
ANC
production
Alkalinity,
S042", NO3-
Ca2+, Mg2+
N/Ap
Rivers
U.S.
Varies,
average
monitoring
data
spanned
50 yrs
(range 27 to
65 yr)
Varies	Increasing alkalinity observed
in 14 of 23 rivers attributed to
recovery from acidification,
agricultural processes and
changing land uses.
Stets etal. (2014)
June 2018
7-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
In the most recent U.S. EPA National Lakes Assessment, based on lake surveys
conducted in 2012, more than 1,000 lakes, ponds, and reservoirs (representing
approximately 100,000 water bodies) were sampled for their water quality and biological
and habitat conditions using generally comparable field and laboratory protocols (U.S.
EPA. 2016h). Inclusion criteria were water bodies >1 hectare, at least 1 m deep, a
minimum of 0.1 hectare open water and minimum water residence time of 1 week.
Chemical indicators in the lake surveys included acidification. This survey updated the
previous National Lakes Assessment (U.S. EPA. 200%). In the earlier National Lakes
Assessment that was conducted in 2007 (U.S. EPA. 200%). ANC and pH were among
the chosen chemical indicators used to assess biological integrity. However, due to
logistical considerations, included sites were restricted to lakes larger than 10 acres
(4 ha). This constraint eliminated from consideration many of the most acid-sensitive and
acid-impacted lakes. In the more recent 2012 lakes survey, the size criterion for inclusion
was decreased to lakes >1 hectare (U.S. EPA. 2016h).
The U.S. EPA also conducted a National Rivers and Streams Assessment (NRSA) during
the period 2008-2009 (U.S. EPA. 2016i). Nearly 2,000 perennial river and stream sites
were sampled, representing nearly 1.2 million miles of stream reach. Stream size ranged
from very small headwaters to very large rivers. The NRSA reported four chemical
stressors: total N, total P, salinity, and acidification along with biological indicators
including benthic microinvertebrates and fish. In the 2012 survey, 97% of lakes were
classified in the least disturbed condition for acidification. During the period 2008-2009
(U.S. EPA. 2016D acidification was found to be a problem in less than 1% of the stream
and river length in the U.S. (U.S. EPA. 2016i). These results emphasize the finding that
acidification impacts are largely confined to relatively small areas of high sensitivity
and/or to relatively small lakes and tributary streams.
The U.S. Geological Survey (USGS) has operated several long-term monitoring efforts,
including the National Ambient Water Quality Assessment (NAWQA) program that was
created in 1991 to assess the nation's water quality in 51 study units defined primarily by
major drainage divides. These units comprise approximately 50% of the conterminous
U.S. Major objectives of the NAWQA program have been to determine the condition of
the nation's streams, rivers, and groundwater; whether these conditions are changing over
time; and how these conditions are affected by natural features and human activities. The
major priority of the NAWQA program since its inception has been the study of
watersheds that have experienced impacts from agriculture and various forms of
development. The location of sites, sampling frequency, and types of measurements all
reflect this priority. The NAWQA data set has also been used to develop predictive
models for occurrence of macroinvertebrates and fish in U.S. streams and identify the
relative importance of factors, including nutrients, in species distributions (Carlisle and
7-40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Meador. 2007; Carlisle et al.. 2007; Meador and Carlisle. 2007). Additional USGS
monitoring programs have included the Hydrologic Benchmark Network (discontinued in
1997), the New York District of the USGS for the water supply watershed for New York
City (in the Catskill Mountain region), and a monitoring project conducted by USGS in
the Adirondack Mountains at Buck Creek, NY.
Studies by Kaushal et al. (2013) and Stets et al. (2014) assessed changes in, and controls
on, carbonate alkalinity of rivers in the U.S. Alkalinity was defined (in equivalent
concentrations, expressed as j^icq/L) as:
[ALK] = [HCOs"] + [COs2"] + [OH"] - [H+]
Equation 7-1
which is mathematically equivalent to ANC defined as the difference between the
equivalent sum of the base cations and the mineral acid anions (Charles and Christie.
1991). Kaushal et al. (2013) reported trends in alkalinity in 97 rivers and streams in the
eastern U.S. They found significant increases in alkalinity, a product of chemical
weathering, at 62 (64%) of the study sites, with no sites showing significant decreases.
Trends of increasing alkalinity were weakly related to the presence of carbonate lithology
in the watershed, acidic deposition, and topography. The authors interpreted the generally
increasing alkalinity as reflective of human impacts on weathering rates. The role of
temperature in modifying weathering rates was not assessed.
7.1.3.2. Eutrophication
A number of monitoring studies have been conducted or continued in recent years that
have documented the effects of N input on the NOs flux and/or trophic state of fresh
waters. Monitoring studies linked to eutrophication response and published since the
2008 ISA are highlighted in this section. In the most recent National Lakes Assessment,
about one-third of the lakes in the continental U.S. were judged to have excess N,
suggesting that N nutrient pollution is a widespread concern across the country (U.S.
EPA. 2016h). Chemical indicators in the lake surveys included DO, N, and P. Benthic
macroinvertebrates, chlorophyll a, and zooplankton were the biological indicators
assessed by the sampling protocols. Algal toxin (microcystin) and cyanobacteria were
also evaluated in the water bodies. Lakes that had high levels of N were 1.6 times as
likely to have a degraded benthic macroinvertebrate community. Increases in both
cyanobacteria (8.3% increase) and microcystin toxin (9.5% increase) were reported in the
most recent lake sampling compared to the previous National Lakes Survey conducted in
2007 (U.S. EPA. 2016h. 2009b). In the NRSA, 41% of the nation's river and stream
miles were rated poor for N compared to least-disturbed reference conditions (U.S. EPA.
June 2018
7-41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
20161). Streams in the Plains and Lowlands were most impacted by N, followed by the
mountainous areas east of the Mississippi River (Eastern Highlands).
Since the 2008 ISA, surface water chemistry data from long-term monitoring by federal,
state, and local agencies as well as university research groups and nonprofits has been
combined into publically available metadatabases. The Lake Multi-Scaled Geospatial and
Temporal Database of the Northeast Lakes of the U.S. (LAGOS-NE) is an integrated
database that includes lake surface water chemistry data from a 17-state region (Soranno
et al.. 2015). A Georeferenced Lake Nutrient Chemistry (GLNC) database containing N
and P water chemistry from 1964-2015 from 3,602 western U.S. mountain lakes has
combined data from the National Park Service (NPS), U.S. Fish and Wildlife Service
(USFS), the U.S. EPA National Lakes Assessments, and academic researchers for
assessing nutrient deposition effects in the region (Williams et al.. 2017b). These
databases have been used to assess temporal and spatial trends of lake nutrient chemistry
in response to environmental stressors including atmospheric deposition.
Research published since the 2008 ISA and described in detail in Appendix 9.1.1.4
provided evidence suggesting that predominantly dry deposition of fine (<10 |_im) and
coarse (<100 |_im) particulates containing P "dust", play a role in the enrichment effects
of N deposition to fresh waters and their catchments. Among the large natural P emission
sources are soils, vegetation, and biomass combustion ash. Other notable sources include
industry, agriculture, and mining. Although P is not a criteria pollutant, inputs of P may
contribute to eutrophication and affect shifts in lake trophic status from P to N limitation
or to colimitation. Data from the U.S. EPA National Lakes surveys and National Rivers
and Streams surveys were analyzed by Stoddard et al. (2016) to determine whether total
P (TP) concentrations changed between 2000 and 2014. They found increases in TP
continentally and especially at sites that exhibited low disturbance (Appendix 9.1.1.2). In
addition, they observed that TN was strongly correlated with TP in lakes and streams on a
national scale. Although the authors determined that TP was increasing at "minimally
disturbed sites," they observed that TN was not increasing at those sites. A 5-year,
Community Atmospheric Model (CAM4) simulation by Brahnev et al. (2015) suggested
that P deposition may play a large role in alpine lake trophic status and that TP deposition
may have increased globally by 1.4 times the preindustrial deposition rate. A global-scale
analysis by Tipping et al. (2014) suggested that oligotrophic lakes are most likely to
experience effects of atmospheric P deposition, with implications for changes in
productivity in response to anthropogenically emitted N.
Regional monitoring data have been used to infer the N saturation status of watersheds in
forested ecosystems based on NO3 leaching to surface waters. Yanai et al. (2013)
conducted a detailed time-series analysis of precipitation and stream chemistry data
7-42
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
(1965-2007) to examine the N budget of a forested reference watershed at HBEF in New
Hampshire. They found that decreases in stream NOs preceded decreases in atmospheric
N deposition. Eshleman et al. (2013) evaluated changes in the concentrations of NO3 in
streams impacted by atmospheric N deposition in the Appalachian Mountains during the
period 1986-2009. Long-term monitoring of nutrient dynamics has also been conducted
in the western U.S. Mastetal. (2014) measured long-term changes in stream NO3
concentration over three decades at the Loch Vale watershed in Rocky Mountain
National Park. During the last 15 years, surface waters in the Green Lakes Valley at the
Niwot Ridge Long-Term Ecological Research site in Colorado have been subjects of a
wide variety of monitoring and process studies that focused on surface waters as
integrators of changes in environmental conditions and the response to external stressors,
including atmospheric deposition and climate change (Miller and Mcknight. 2015; Elser
et al.. 2009b; Elser et al.. 2009a; Gardner et al.. 2008).
7.1.4. Models
Models used to assess the effects of N and S deposition on U.S. ecosystems were
described in the 2008 ISA (Annex A) and Appendix 4 of this ISA. The most frequently
used ecosystem models for aquatic systems situated in small watersheds have included
the Model of Acidification of Groundwater in Catchments (MAGIC) and the
Photosynthesis and Evapotranspiration-Biogeochemical (PnET/BGC) model lYU.S. EPA.
2008a'); Appendix 4.51. ForSAFE Wallman et al. (2005) has also been applied widely,
especially in Europe, and has been linked with a terrestrial plant biodiversity model,
VEG, and recently applied at several locations in the U.S. lYPhelan et al.. 2016;
McDonnell et al.. 2014a); Appendix 4.5.1.41. The Very Simple Dynamic (VSD) soil
acidification model is used in Europe to simulate acidification effects in soils when
observed data are sparse (Appendix 4.5.1.3'). It has not been used widely in the U.S.
Three other models, SPARROW, Watershed Assessment Tool for Evaluating Reduction
Scenarios for Nitrogen (WATERS-N), and the Surface Water Assessment Tool (SWAT)
have been used to evaluate N loading to large river systems. Another model that has been
applied to analysis of nutrient enrichment in aquatic systems is AQUATOX (Carleton et
al.. 2009). which simulates nutrient dynamics and effects on aquatic biota. Such models
were earlier summarized in Table A-8 in the 2008 ISA (U.S. EPA. 2008a).
In the 2008 secondary NAAQS review for oxides of N and S, an aquatic acidification
index (AAI) was developed to relate (1) atmospheric concentrations of SOx and
NOy + NHy to N and S deposition levels using transference ratios (Appendix 2) and
(2) to relate deposition to ANC values, using a modified Steady-State Water Chemistry
(SSWC) model (Appendix 4) and water chemistry data for over 6,000 sites in the U.S.
June 2018
7-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
The ANC values were grouped by site into ecoregions and evaluated by considering the
distribution of predicted ANC values (Scheffe ct al.. 2014).
7.1.4.1.	Updates to Key Previously Identified Models
Zanchi etal. (2016) developed an approach to include lateral flow in ForSAFE. Results
were assessed by comparison with research values at the Vindela Research forest in
northern Sweden. Simulation of both saturated and unsaturated soil zones improved
agreement between measured and modeled water flows. This model improvement will
likely enhance the ability to simulate the export of elements from soil to drainage water
using ForSAFE. In 2011, The SPARROW modeling group published a new set of
regional models as a special issue (Preston et al.. 201IV Other models have been used for
assessing nutrient loading to freshwater systems. SWAT has been adopted as part of
U.S. EPA's Better Assessment Science Integrating Point and Nonpoint Sources
(BASINS) platform in support of TMDL model development (Gassman et al.. 2007).
7.1.4.2.	New Models (Published since 2008)
Several new freshwater acidification or eutrophication models have been developed and
published since 2008. The Watershed Analysis Risk Management Framework (WARMF)
is a model for evaluating point and nonpoint pollution, including N fate and transport
(Davvani et al.. 2013; Herr and Chen. 2012). A new national water quality modeling
system (Hydrologic and Water Quality System [HAWQS]) is under development by
Texas A&M University and U.S. Department of Agriculture (USDA) for U.S. EPA's
Office of Water (https: //epahawq s .tamu. edu/). The model is intended to assist resource
managers and policy makers in evaluating the effectiveness of water pollution control
efforts. HAWQS will support application of SWAT and SPARROW and will include
simulations of nutrients and other pollutants. Table 7-4 summarizes recent process-based
model estimates of surface water acidification and chemical recovery in the U.S.
June 2018
7-44
DRAFT: Do Not Cite or Quote

-------
Table 7-4 Recent process-based model estimates of surface water acidification
and chemical recovery in the U.S.
Nutrient
Model Enrichment
Acidification
Type of
Ecosystem Region
Model Application and
Findings
Publication
PnET-
BGC
X
30 Southern
watersheds in Appalachian
Great Smoky Mtns.
Mts. NP
Stream recovery has been
limited.
Fakhraei et al.
(2016)
PnET-
BGC
X
44 Adirondack
representative Mtns.
watersheds
Larger historical lake
acidification in lakes having
lower ambient ANC.
Zhai et al. (2008)
MAGIC	X 66 stream Southern All modeled streams had
watersheds Blue Ridge preindustrial ANC
Mtns.	>30 peq/L. Median stream
lost about 25 peq/L of ANC
between 1860 and 2005.
Of 128 acid-impaired lakes Fakhraei et al.
that were modeled, 97 had (2014)
ambient ANC below target
of 20 peq/L and 83 were
below target of 11 peq/L.
TMDL corresponding to a
moderate emissions
control scenario (60%
decrease in S deposition)
was 7.9 meq/m2/yr.
Simulations under	Pourmokhtarian
changing climate reflected et al. (2012)
later snowpack
development, earlier spring
snowmelt, greater ET, and
slight increase in water
yield. Net soil
mineralization and
nitrification caused
simulated soil and water
acidification.
Hindcast and forecast Tominaaa et al
projections were	(2010)
qualitatively similar, but
temporal patterns of
simulated change in
chemistry differed
substantially among
models.
Sullivan et al
(2011b)
PnET-	X
BGC
128	Adirondack
acid-impaired Mtns.
lakes
PnET-	X Streams, HBEF, NH
BGC	watershed
PnET-	Inter-model HBEF, NH
BGC,	comparison
SAFE,
VSD,
MAGIC
June 2018
7-45
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Table 7-4 (Continued): Recent process based model estimates of surface water
acidification and chemical recovery in the U.S.
Nutrient
Model Enrichment Acidification
Type of	Model Application and
Ecosystem Region	Findings
Publication
PnET-
BGC
X Three	Adirondack Predicted ANC recovery
multipollutant Mtns.	closely related to
scenarios	percentage of watershed in
conifers, elevation, and
lake area.
Wu and Driscoll
(2009)
ANC = acid neutralizing capacity; ET = evapotranspiration; HBEF = Hubbard Brook Experimental Forest; L = liter; m = meter;
|jeq = microequivalent; meq = milliequivalent; MAGIC = Model of Acidification of Groundwater in Catchments;
PnET-BGC = Photosynthesis and Evapo Transpiration-Biogeochemical; S = sulfur; SAFE = Soil Acidification in Forest
Ecosystems; TMDL = total maximum daily load; VSD = Very Simple Dynamic; yr = year.
7.1.5. National Scale Sensitivity and Response
The 2008 ISA documented that by the end of the 1980s, the regions of the U.S. with
many acid-sensitive waters and ecosystems were well recognized. These acid-sensitive
ecosystems are mostly located in upland mountainous terrain in the eastern and western
U.S. and are underlain by bedrock that is resistant to weathering, such as granite or
quartzite sandstone. However, a similar map for areas sensitive to the eutrophication
effects of N were not available. There is strong evidence demonstrating that
biogeochemical sensitivity to deposition-driven eutrophication and acidification is the
result of historical loading, geologic/soil conditions (e.g., mineral weathering and S
adsorption), and nonanthropogenic sources of N and S loading to the system. There is no
single deposition level applicable to all ecosystems in the U.S. that can describe the onset
of eutrophication or acidification. There are new publications that address recovery of
freshwater ecosystems at either the national scale or in specific regions. National scale
sensitivity of freshwater systems to N-nutrient effects is discussed in Appendix 9.1.1.2.
7.1.5.1. Chemical Recovery
Chemical recovery of previously acidified surface water is required to support biological
recovery. An aquatic ecosystem in chemical recovery will have trends in water quality
indicators (NO3 ,SC>42~, pH, ANC, inorganic monomeric Al) towards inferred
preindustrial values (Chapter 1). Preindustrial water quality indicator values are inferred
from models, paleolimnology samples, or historical samples of biological communities.
Preindustrial water quality varied across the conterminous U.S. in response to variation in
climate, geology, and biological communities. In general, biological recovery has lagged
June 2018
7-46
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
behind chemical recovery in previously acid-impacted and more recently recovering
surface waters (Appendix S).
Surface water NOs is a chemical indicator for both eutrophication and acidification.
Several studies since the 2008 ISA have documented decreased surface water NO3
concentration attributed to decreases in atmospheric deposition. Eshleman et al. (2013)
evaluated changes in the long-term concentration of NO3 in surface waters impacted by
atmospheric N deposition in the Appalachian Mountains during the period 1986-2009.
Regional NOx emissions progressively decreased during the study period. Decreases in
the annual surface water NO3 concentration and the NO3 yield were observed over the
study period, which corresponded to generally comparable declines in annual wet N
deposition. Eshleman and Sabo (2016) evaluated changes in NO3 concentration over
time in tributary streams of the upper Potomac River. The basin-wide decrease, based on
results at the Washington, D.C. station was -0.023 mg N/L/year over 26 years (total
change of-0.59 mg N/L). They attributed observed decreases in discharge-weighted
annual mean NO3 concentrations across the basin (mean decrease was 37%) largely to
decreases in atmospheric N deposition. Mast et al. (2014) measured long-term changes in
stream NO3 concentration over three decades at the Loch Vale watershed in Rocky
Mountain National Park. The concentrations of NO3 in stream water increased during
the early 1990s, peaked in the mid 2000s, and then declined by more than 40%. The
recent decreases in stream NO3 corresponded with decreases in NOx emissions and N
deposition. As was found by Eshleman et al. (2013) in Maryland, similarities in the
timing and magnitude of NO3 concentrations in stream water and N deposition
suggested that stream chemistry is responding to changes in atmospheric deposition of N.
However, the response was complicated in this case by a drought in the early 2000s that
enhanced N export for several years. In the Adirondack region concentration of NO3
declined at variable rates in 33 of the 48 study lakes since 1992 (Driscoll et al.. 2016).
Strock et al. (2014) analyzed recent trends in lake chemistry using long-term data from
lakes in the Adirondack Mountains and New England. Lake NO3 concentration showed
no trend prior to 2000. During the 2000s, the wet deposition of NO3 declined more than
50%, and lake NO3 concentration declined subsequent to 2000 at a rate of
-0.05 (ieq/L/yr. TN declined in surface waters at a rate of 1.1% per year from 1990 to
2011 in an analysis of lake data from 17 states in the Midwest and northeastern U.S.
(Oliver etal.. 2017). Random forest analysis of the data showed atmospheric deposition
was the top driver of observed declines in TN.
Evidence of chemical recovery from acidification has been provided by monitoring
efforts spanning several decades (Appendix 7.1.3) and the use of models to hindcast and
forecast the acid-base chemistry and N response in soils and surface waters
(Appendix 7.1.4). Model projections of past and future changes in surface water
7-47
DRAFT: Do Not Cite or Quote

-------
1
2
chemistry in response to changes in acidic deposition using the MAGIC and PnET-BGC
models are summarized in Table 7-5.
Table 7-5 Model projections of surface water sulfate and associated acid
neutralizing capacity, shown as changes between dates, for
Adirondack and Shenandoah streams.
Region
Water
Bodies
Dates
Model
Pollution
Scenario
Change in
Median
Surface
Water SO42"
(jeq/L
Change in
Median
Surface
Water ANC
(jeq/L
Reference
Hindcasts
Adirondacks,
NY
38 lakes
1850 to
2003
PnET-
BGC

+72.9
-39.9
Zhai et al. (2008)
Adirondacks,
NY
37 lakes
1850 to
1984
PnET-
BGC

+ 107
-77.8
Chen et al.
(2005a)
Adirondacks,
NY
44 potentially
acid-sensitive
lakes
1850 to
1990
MAGIC

+77.8
-38.3
Sullivan et al.
- C9nnfiai
PnET-
BGC

+57.3
-29.5

Adirondacks,
NY
141 TMDL
lakes
1850 to
2010
PnET-
BGC

+65.2
-39.4
Fakhraei et al.
(2014)
Forward projections
Shenandoah
NP, VA
Five streams
on siliciclastic
bedrock
1990 to
2040
MAGIC
Constant
deposition
+ 13
-11.6
Sullivan et al.
(2008)



Mild reduction
-21
+6.2





Medium
reduction
-23
+7.2





Strong
reduction
-40
+24.2





Very strong
reduction
-44
+27.2

June 2018
7-48
DRAFT: Do Not Cite or Quote

-------
Table 7-5 (Continued): Model projections of surface water sulfate and associated
acid neutralizing capacity, shown as changes between
dates, for Adirondack and Shenandoah streams.
Region
Water
Bodies
Dates
Model
Pollution
Scenario
Change in
Median
Surface
Water SO42"
(jeq/L
Change in
Median
Surface
Water ANC
(jeq/L
Reference
Shenandoah
Four streams
1990 to
MAGIC
Constant
+22
-8
Sullivan et al.
NP, VA
on granitic
bedrock
2040

deposition


(2008)



Mild reduction
+ 11
-5





Medium
+ 11
-5





reduction







Strong
reduction
+3
-2





Very strong
reduction
+2
-2

Shenandoah
Five streams
1990 to
MAGIC
Constant
+33
-5
Sullivan et al.
NP, VA
on basaltic
bedrock
2040

deposition


(2008)



Mild reduction
+ 12
0





Medium
+ 11
+ 1





reduction







Strong
reduction
-4
+5





Very strong
reduction
-9
+6

Adirondacks,
44 potentially
1990 to
MAGIC
Current and
-42.4
+5.89
Sullivan et al.
NY
acid-sensitive
lakes
2050

expected
controls


(2006a)




Moderate
-58.9
+ 18.6





emissions







controls



Aggressive
emissions
controls
-64.6
+22.6
June 2018
7-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Table 7-5 (Continued): Model projections of surface water sulfate and associated
acid neutralizing capacity, shown as changes between
dates, for Adirondack and Shenandoah streams.
Change in Change in
Median Median
Surface Surface
Region
Water
Bodies
Pollution Water S042" Water ANC
Dates Model Scenario	peq/L	(jeq/L
Reference
Adirondacks, 44 potentially 1990 to MAGIC Current and
NY	acid-sensitive 2050	expected
expected
controls
-18
-3.7 Sullivan et al
(2006a)
lakes
Moderate
emissions
controls
-32.2
+ 1.8
Aggressive
emissions
controls
-38.3
+9.3
ANC = acid neutralizing capacity; L = liter; |jeq = microequivalent; MAGIC = Model of Acidification of Groundwater in Catchments;
PnET-BGC = Photosynthesis and Evapo Transpiration-Biogeochemical; S042" = sulfate.
Source: U.S. EPA (2008a).
As reported in the 2008 ISA, lake and stream ANC values decreased throughout much of
the 20th century in a large number of acid-sensitive lakes and streams throughout the
eastern U.S. This effect has been well documented in monitoring programs,
paleolimnological studies, and model simulations. Since about 1990, the ANC values of
many previously acidified lakes and streams have shown some increases, but such increases
have been relatively modest in most cases. Kahl et al. (2004) summarized 20 years of
chemical monitoring data from regional U.S. EPA programs in the northern and eastern
U.S. Surface water chemistry improved over the 20-year period in many lakes and streams;
moderate, but significant increases in ANC and pH were observed. Stets et al. (2014) found
widespread increasing alkalinity concentrations at 14 of 23 sites across the conterminous
U.S. over the last half of the 20th century and the early 21st century. Results showing
decreases in NOs and SO42 concentrations and cation: alkalinity ratios at many sites were
consistent with recovery from water acidification. Agricultural lime contributed alkalinity
at some locations. These findings are important because large rivers constitute important
pathways for transport of nutrients and other constituents to coastal waters and may
influence coastal acidification and Ca supply.
Trends in surface water chemistry, including evidence for chemical recovery, are reported
in several acid-sensitive regions of the U.S. where long-term monitoring programs have
been in place several decades. These include the Adirondacks and the northeastern U.S.(see
Appendix 16). In some portions of the Appalachian Mountains, chemical recovery of
June 2018
7-50
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
surface waters has not always followed decreases in deposition due to high S absorption in
watershed soils (Appendix 16 Smoky Mountains Case Study).
A long-term delay in surface water recovery from past acidification has been noted by
Likens et al. (1996) at HBEF. This delay was attributed mainly to changes in base cation
concentration. Similarly, based on an analysis of Adirondack lakes, Lawrence et al. (2013)
concluded that chemical recovery processes in response to decreases in acidic deposition
are complex and require evaluation of multiple chemical metrics in addition to ANC,
including DOC, inorganic Al, and BCS. As an example of the complexity, the lower
threshold of ANC required to protect brown trout in Norway against acidification damage
was earlier estimated to be 20 (ieq/L (Lien et al.. 1996).This critical ANC limit may be
lower in humic lakes (~8 j^ieq/L) because about one-third of the organic acid anions may act
essentially as strong acid anions (Lvdersen et al.. 2004; Driscoll et al.. 1994). Inclusion of
strong organic acids in the ANC calculation can be useful because the relationships
between pH, inorganic Al, and ANC may have changed in response to increased
concentrations of DOC and TOC. This is particularly true in lakes having higher DOC and
TOC compared with clear water lakes. Hesthagen et al. (2008) estimated that threshold
ANC values to prevent damage to brown trout had increased by 1995 to 48 (ieq/L. The
authors suggested that the higher ANC threshold for 1995 was attributable to lower pH and
higher inorganic Al at a given ANC level.
Adirondacks
At the time of the 2008 ISA, several studies reported chemical recovery from acidification
in Adirondacks lakes (Driscoll et al.. 2007a; Momen et al.. 2006; Driscoll et al.. 2003c)
based on long-term monitoring efforts in the region (Appendix 7.1.3.1). Driscoll et al.
(2003c') evaluated changes from 1982 to 2000 in the original 16 Adirondack LTM lakes
and from 1992 to 2000 in the complete set of 48 Adirondack LTM lakes. They found that
nearly all study lakes showed marked decreases in SOr concentration and several lakes
showed declines in NOs concentration. They found that 7 of the 16 original monitoring
lakes showed statistically significant increases in ANC, with a mean rate of increase of
0.78 (j,eq/L/year. In a study of 30 of the 48 lakes studied by Driscoll et al. (2003b). pH
increased in 25 lakes and ANC increased in 12 of the 30 lakes (Momen et al.. 2006).
Concentrations of dissolved NO3 were inversely correlated with concentrations of
chlorophyll a in 11 lakes. Chlorophyll a increased in concentration in 9 lakes. The increase
in pH observed in most of these lakes may have stimulated productivity so that N
assimilation by plankton increased, indicating recovery from acidification to
eutrophication, rather than to preindustrial conditions.
7-51
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Monitoring data since the 2008 ISA continues to show chemical recovery in the
Adirondack region. As described in the Adirondacks case study (Appendix 16). significant
decreases in lake SOr concentrations (-2.14 (imol/L/year) corresponding to significant
declines in total S deposition were observed in monitoring data from 1984 to 2010
(Mitchell et al.. 2013). Surface water ANC and pH recovery has been documented by
Lawrence et al. (2013) and Lawrence et al. (2011) in studies of Adirondack streams and
lakes. On average, the pH increased by only 0.28 pH units and ANC by 13 (j,eq/L in
12 streams that were monitored over a period of 23 years. During that period, there was an
approximate 50% reduction in atmospheric deposition of S in the Adirondack region. The
percentage of Adirondack lakes that were acidic decreased from an estimated 15.5 to 8.3%
using data from the TIME monitoring program (Waller et al.. 2012). Decreases in lake
water SO42 . and to a lesser extent NO3 . concentrations generally were accompanied by
increases in lake ANC. Driscoll et al. (2016) observed increases in ANC and pH and
marked decreases in dissolved inorganic Al in 45 of 48 study lakes in the Adirondack
region from 1982 to 2015, corresponding to decreases in acidifying deposition. Changes
were most pronounced in the acid-sensitive thin-till drainage lakes. In another Adirondack
long-term study that focused on 43 lakes sampled by the Adirondack Lakes Survey
Corporation during three time periods (1984-1987, 1994-2005, and 2008-2012), the
average concentration of inorganic Al decreased from 2.2 (.iM (a level above that generally
considered toxic to brook trout, 2 (.iM) to 0.66 (.iM (Baldigo et al.. 2016).
Interpretation of long-term trends in Adirondack surface water chemistry, as summarized
above, has also been augmented by results of repeated surveys of the chemistry of lakes
initially surveyed several decades previously. Warbv et al. (2008) resurveyed 113 lakes
throughout the Northeast in 2001 that had previously been sampled by the U.S. EPA in
1986 to assess chemical recovery from lake acidification. Decreases in total Al and
inorganic Al concentrations were widespread, and were largest in the Adirondack
Mountains (-2.59 to -4.60 (imol/L). In 2001, only 7 study lakes (representing 130 lakes in
the population) had inorganic Al concentrations above a toxic limit of 2 |imol/L. compared
with 20 sampled lakes (representing 449 lakes in the population) in 1986. Thus, it was
estimated that in 2001 more than 300 northeastern lakes no longer had summer inorganic
Al concentrations at levels considered harmful to aquatic biota. Michelena et al. (2016)
reported changes in the water chemistry of 30 Adirondack lakes in response to reductions
in acidic deposition from 1994 to 2012. The water quality of the study lakes generally
improved during the study period, but the responses were sporadic and complex. Lake pH
values increased until about 2002 and then fluctuated. Inorganic Al concentrations
generally decreased throughout the period of record. During the early years of monitoring,
the average pH increased dramatically from about 5.5 to 6.0. This increase was followed by
a period of reacidification for about 5 years, followed by another period of increased pH.
7-52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
This 5-year cycle was then repeated to a muted extent, even though acidic deposition
continued to decline.
Chemical recovery from surface water acidification in the Adirondack Mountains has been
accompanied by increasing concentrations of DOC, including organic acids, which have
complicated and restricted recovery from acidification (Section DOC). Long-term
monitoring of 48 Adirondack lakes by Driscoll et al. (2016) showed patterns of increasing
DOC in 29 study lakes concurrent with decreases in acidic deposition. Lawrence et al.
(2013) evaluated long-term changes in DOC and BCS (which reflects the calculated ANC
and includes an adjustment for strong organic acid anions). Increases in DOC concentration
from 1994 to 2011 contributed to organic complexation of Al. This reduced the fraction of
monomeric Al that was in the inorganic (and potentially toxic) form from 57% in 1994 to
23% in 2011. Thus, the higher DOC found during latter sampling increased the organic
acidity of the lake water and limited ANC recovery, but at the same time it decreased the
toxicity of dissolved Al to aquatic biota. Similarly, Strock et al. (2014) reported trends in
recovery from toxic Al levels caused by acidification of surface waters. They analyzed
recent trends in lake chemistry using long-term data from lakes in the Adirondack
Mountains and New England. During the 2000s, the wet deposition of NO3 declined more
than 50%. The lake NOs concentration, which showed no trend prior to 2000, declined
subsequent to 2000 at a rate of-0.05 (j,eq/L/year. There was a shift to nontoxic (organic)
forms of Al. Despite the Al recovery, both the ANC and pH in the study lakes failed to
show evidence of appreciable chemical recovery; rather, the lakes continued to exhibit
variable trends in ANC and pH.
The PnET-BGC model has been applied in the Adirondack region to assess chemical
recovery. Model results from Fakhraei et al. (2014) suggested that ambient ANC values
were below the target value of 20 (j,eq/L in 97 of the 128 lakes that were judged to be acid
impaired under Section 303(d) of the Clean Water Act (CWA); 83 lakes had ANC below
the target value of 11 (j,eq/L. A moderate emissions control scenario (60% decrease from
the ambient atmospheric S deposition) was projected to recover the ANC of lakes by 0.18
and 0.05 (ieq/L/year on average during the periods 2022 to 2050 and 2050 to 2200,
respectively. Model results suggested that controlling S deposition to Adirondack lake
watersheds was more effective as a means to recover acidic lakes than was controlling N
deposition in this region. Wu and Driscoll (2009) applied the PnET-BGC model to predict
the responses of lake ANC during the period 2001 to 2050. Based on three multipollutant
scenarios, predicted ANC recovery was associated with higher elevation, smaller lake area,
and higher percentage of watershed area in coniferous vegetation. The variables lake depth
and the square of the lake elevation explained 40% of the variation in predicted lake water
ANC.
7-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Northeast
The Northeast U.S. also has several decades of monitoring data, especially for HBEF
(Appendix 7.1.3.1). Strock et al. (2016) documented decreases in water transparency and
increases in DOC in lakes in Acadia National Park in Maine over a period of 14 years
(1995-2008). Larger transparency decreases were noted in clear water lakes (-0.3 m/year)
compared with the brown water lakes (-0.1 m/year). In the Catskill Mountains, NY,
23 years of stream water chemistry data showed significant decreasing SO42 (mean trend
of-2.5 (ieq/L/year), while no significant trends were observed for NO3 (Mchalc et al..
2017). A decreasing trend in inorganic Al and increasing trends in pH and ANC were
evident under both low- and high-flow conditions. Fuss et al. (2015) examined long-term
trends in soil solutions and surface waters from the early 1980s to 2011 at the HBEF in
New Hampshire. They found that rates of annual average ANC increase during the period
1982-2011 were similar to what was observed during snowmelt, although the ANC during
snowmelt was 10 j^icq/L lower than annual averages. In a summary of 20 years of data at
Bear Brook, Norton et al. (2010) observed decreases in pH, base cation concentrations, and
ANC, and increases in inorganic Al concentration in the East Bear reference watershed.
These observations indicate that under ambient deposition conditions acidification
continues even though S deposition and stream SO42 concentrations have declined. Laudon
and Norton (2010) analyzed 212 hydrological episodes in Bear Brook using the ANC
Dilution Model (ADM) of Laudon and Bishop (1999). The results showed that 18 years of
experimental addition of N and S to the West Bear Brook watershed had not altered the
most important natural causes of episodic stream acidification: base cation dilution,
processes associated with deposition of marine sea salts, and contributions of organic
acidity during rain and snowmelt events. The results further indicated that the contributions
of S042 to the observed ANC decreases in West Bear Brook during episodes increased
steadily since the beginning of experimental watershed treatment. In contrast, the role of
N03 in episodic acidification events remained relatively constant after an early increase.
Model stimulations at HBEF using the PnET-BGC model showed later development of
snowpack, earlier snowmelt, higher evapotranspiration, and increased water yield expected
with current and projected trends of increasing temperature will increase net soil N
mineralization and nitrification (Pourmokhtarian et al.. 2012). This could contribute to
acidification of soil and stream water. In a model study comparison, Tominaga et al. (2010)
noted that hindcast (1850-1992) and forecast (2005-2100) projections were qualitatively
similar across MAGIC, PnET-BGC, SAFE, and VSD watershed acidification models at
HBEF, although projected stream ANC and soil base saturation differed substantially
through time.
7-54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
West
Surface water acidification, and presumably also chemical recovery, have been limited in
the western U.S. Although acidification sensitivity has been shown to be high at many
locations, acidic deposition levels have also mostly been lower. Mastetal. (2011)
examined trends in precipitation chemistry and other factors that influence long-term
changes in chemistry of high-elevation lakes in three Colorado wilderness areas during the
years 1985 through 2008. Sulfate concentrations in precipitation decreased at rates of-0.15
to -0.55 (j,eq/L/year at 10 monitoring stations in Colorado. In lakes where SO42 was
primarily derived from atmospheric sources, the lake SO42 concentrations decreased by
-0.12 to -0.27 (j,eq/L/year. In lakes where SO42 likely originated primarily from watershed
weathering, the SO42 concentrations in lake water increased, rather than decreased, from
1985 to 2008.
Appalachians
Not all acid-sensitive regions of the U.S. have shown improvements in surface water
quality with decreasing deposition trends. Monitoring data from the Great Smoky
Mountains NP (Appendix 16) indicate that the high S absorption in watershed soils in this
region delay recovery from previous stream acidification (Cai et al.. 2010). The majority
(about 61%) of the net SO42 entering the study watershed was retained. However, during
large precipitation events, SO42 in wet deposition moved more directly and rapidly to
streams, contributing to episodic stream acidification. Cai et al. (2010) concluded that base
cation depletion from soils could limit chemical recovery from acidification of streams in
the study watershed. Stream chemistry from 1993-2014 at 42 monitoring sites in the park
did not show substantial changes over the recent period of long-term monitoring (Fakhraei
et al.. 2016). An empirical modeling study by Robinson et al. (2008) of baseflow water
chemistry at 90 streams in Great Smoky Mountains National Park during the years
1993-2002 indicated significant decreasing trends in stream pH and SO42 at lower
elevation sites overtime, but no long-term trends in stream NO3 or ANC. An estimated
30% of the sample sites were simulated to decrease pH to less than 6 within 10 years. The
models explained 71% of the variability in pH and 86% of the variability in ANC. Soil
retention of SO42 and assimilation of NO3 into plants were important biogeochemical
processes regulating stream chemistry. The stream pH at base flow has not yet shown signs
of recovery in response to recent decreases in acidic deposition. The model developed for
elevations between 305 and 1,070 m predicted that median pH values will decrease below 6
within 34 years if the observed statistical trends continue. The state of Tennessee currently
lists streams within the park with mean pH below 6 as impaired under the Clean Water Act
7-55
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Section 303d (Appendix 16. Southeast case study), so the model by Robinson et al. (2008)
indicated biologically significant ongoing acidification.
Singh et al. (2016) assessed DOC trends in a forested stream at the Coweeta Hydrologic
Laboratory in the southeastern U.S. over a period of 25 years. Concentrations and fluxes of
DOC decreased by 34 and 56%, respectively, between 1988 and 2001, corresponding with
the stream acidification phase. During the period 1997 to 2012, DOC concentration
increased in association with increases in precipitation and the number and intensity of
short-duration storms during the early part of the growing season.
Several model applications have been conducted in the southern Appalachian Mountains to
simulate stream chemistry and critical or target loads. These studies have employed the
MAGIC, PnET-BGC, and SSWC models. Results of these studies have reinforced the
widespread acid sensitivity of this region and the importance of S adsorption and base
cation depletion to stream responses to changes in levels of acidic deposition. Critical and
target loads (Chapter 1.2.2.3) for resource recovery/protection were quantified | (Fakhraci et
al.. 2016; McDonnell et al.. 2014b; McDonnell et al.. 2013; McDonnell et al.. 2012;
Sullivan et al.. 201 lc; Mcdonnell et al.. 2010); Appendix 8.5.41. An important outcome of
critical load modeling studies in the Appalachian Mountains is the suggestion that complete
stream acid base chemistry recovery may not be possible (Appendix 8.5.4). Sullivan et al.
(201 lb) applied the MAGIC model to estimate the S target load for protecting aquatic
resources in 66 stream watersheds in the Southern Blue Ridge province of the Appalachian
Mountains at different points in the future. For some of the modeled sites, if S deposition
was decreased to zero and maintained at that level throughout the simulation, one or more
of the selected critical ANC levels (0, 20, 50, 100 j^ieq/L) could not be achieved by 2100.
This was likely largely due to the simulation result suggesting that many of the streams did
not exhibit such a high ANC during preindustrial times, in the absence of acidic deposition.
For other sites with large watershed acid buffering capacity, even very high-sustained S
deposition would not reduce stream ANC below thresholds associated with biological
harm.
Monitoring studies suggested that the rate of stream ANC recovery in this region may be
slow in the Southeast and that base cation depletion in soils contributed in the last two
decades to further aquatic acidification despite reductions in S deposition (Robison et al..
2013; Cai et al.. 2010). Rice et al. (2014) examined the source-sink behavior of SO42 in
27 unglaciated forested watersheds across a latitudinal gradient from Pennsylvania to
Georgia and found that many of the watersheds still retain SO42 under conditions of
decreased S deposition. The specific years when the watersheds will likely cross over from
retaining to releasing SO42 varied from north to south, with the south generally showing
later cross-over dates. Eshleman et al. (2008) used data from two long-term monitoring
7-56
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
stations in western Maryland to assess the recovery of stream water ANC due to declines in
acidic deposition. Stream water SO42 concentration declined at the two study sites between
1990 and 2005 by about 3 (ieq/L/year in response to a 34% reduction in wet atmospheric S
deposition. However, trends in NO3 concentration were more strongly related to watershed
factors, especially forest disturbance. Although ANC increased throughout the study, the
rate of increase in later years (1996-2005) was only about half as large as the rate of
increase in ANC over the entire study period. This result might suggest a slowing of
chemical recovery from previous acidification (Eshleman et al.. 2008). This supposition
was further substantiated by Robison et al. (2013). who analyzed the chemistry of stream
samples collected quarterly from 1987 to 2011 at 64 sites in the southern Appalachian
Mountains of western Virginia. At most of the study streams, the pH increased over time.
However, at most sites underlain by base-poor bedrock, ANC decreased even though S
deposition decreased. These decreases in stream ANC were associated with depletion of
base cations in watershed soils.
Stream acid-base chemistry from 1999 to 2014 in a group of 40 stream reaches in the
Upper Savage River watershed in western Maryland showed statistically significant
decreases in stream concentrations of SO42 and NO3 that were qualitatively and
quantitatively consistent with decreases in wet S and N deposition (Kline et al.. 2016).
Stream ANC increased by significant amounts in 10-20% of the monitored streams, but the
magnitude of recovery was too small compared with natural variability to detect a regional
ANC recovery. The percentage of streams having ANC <0 j^ieq/L decreased from about 7%
in 1999 to 0 in 2014. The percentages of streams having ANC values less than 50 j^ieq/L
and less than 100 j^ieq/L also decreased markedly between 1999 and 2014. Concentrations
of base cations (Ca, Mg, K) decreased, moderating regional ANC recovery.
7.1.5.2. Critical Loads
A critical load (CL) is a quantitative estimate of exposure to one or more pollutants below
which significant harmful effects on specified sensitive elements of the environment do not
occur according to present knowledge KSprangcr et al.. 2004; Nilsson and Grennfelt. 1988)
(Chapter 1.2.2.3)1. These CLs can be used as early warning signals to indicate likely
ecosystem sensitivity to change in N or S. Empirical CLs are developed from observational
data while steady-state and dynamic models develop relationships between deposition,
water quality measurements and biogeochemistry for watersheds.
For acidification, CLs from empirical data as well as modeling approaches are available for
acid-sensitive regions of the U.S. (Appendix 8.5.3). Generalized empirical estimates are
based on acidification or increased surface water NO3 leaching. Empirical CLs to protect
June 2018
7-57
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
against aquatic acidification in U.S. ecosystems are summarized in Table 8-7. Both
steady-state (Appendix 8.5.4.1.2.1) and dynamic models (Appendix 8.5.4.1.2.2) have been
used to quantify relationships between deposition and biogeochemistry for watersheds in
order to develop CL estimates to protect against acidification or promote recovery of
acid-base chemistry. Recent CL modeling studies in the U.S. to protect against aquatic
acidification are summarized in Table 8-8. Most aquatic CL studies conducted in the U.S.
have used surface water ANC as the principal metric of water quality change in response to
changes in acidic deposition, although ANC should not necessarily be used as the only
environmental predictor of biological harm on which to base CLs. Other potentially useful
variables include water pH, inorganic Al, and BCS. Since the 2008 ISA, dynamic modeling
of CLs in the U.S. to achieve various ANC targets has been focused mostly on the
Adirondack and Appalachian Mountains. The CL can be calculated to represent the
individual or combined deposition load of S and/or N to which a stream and its watershed
could be subjected and still have a surface water ANC within a targeted range.
For nutrient enrichment, diatoms are among the most sensitive aquatic organisms, thereby
providing a basis for assessing aquatic ecosystem protection against nutrient enrichment
across ecosystems. The lake water NOs concentration has been identified as a useful
chemical criterion indicative of biological change in the diatom community. Recently,
Williams et al. (2017b) used phytoplankton biomass N to P limitation shifts as the basis for
CL calculations. Critical loads (Chapter 1.2.2.3) for nutrient enrichment are described in
Appendix 9^5 and summarized in Table 9-4.
7.1.6. Water Quality Criteria
The term "water quality criteria" is used in two sections of the Clean Water Act:
Section 304(a)(1) and Section 303(c)(2). The term has a different impact in each section. In
Section 304, the term refers to a scientific assessment of ecological and human health
effects that U.S. EPA recommends to states and authorized tribes for establishing water
quality standards that ultimately provide a basis for controlling discharges or releases of
pollutants or related parameters. Ambient water quality criteria associated with specific
water body uses, when adopted as state or tribal water quality standards under Section 303
of the CWA, define the level of a pollutant (or, in the case of nutrients, a condition)
necessary to protect designated uses in specific water bodies.
U.S. EPA develops Water Quality Criteria (WQC) using existing scientific knowledge to
determine when water is unsafe for people and wildlife. These criteria are developed as
recommendations. State and tribal governments may adopt these criteria or use them as
guidance in developing their own. U.S. EPA bases aquatic life criteria on how much of a
June 2018
7-58
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
chemical can be present in surface water before it is likely to harm plant and animal life.
WQC are determined to protect aquatic life, biology, human health, microbial/recreational,
and sediment condition. For aquatic life, the criteria are designed to protect both freshwater
and marine organisms from short-term and long-term exposure to pollutants. Biological
criteria indicate the health of water bodies based on how many and what kinds of organisms
are present. More details about the scientific basis for these criteria may be found at
https://www.epa.gov/wqc/basic-information-water-qualitv-criteria.
The U.S. EPA is working with the states to develop numeric nutrient criteria to better
define levels of N and P that affect U.S. waters. The numeric values include both causative
(N and P) and response (chlorophyll a, turbidity) variables. The U.S. EPA's National
Nutrient Program recommended nutrient criteria for rivers and streams in 14 ecoregions of
the U.S. (based on Omernik Level III ecoregions) to use as starting points for states to
develop their own criteria (U.S. EPA. 1998b).
WQC indicators related to N are available for 10 states to date (Table 7-6) including
Oregon, California, Arizona, Colorado, Montana, Utah, and Mississippi which have
numeric nutrient criteria. These criteria may include a variety of N species and chlorophyll
a (Table 7-6). For Washington and Louisiana, which lack explicit numeric nutrient criteria,
U.S. EPA aggregate Level III ecoregion nutrient criteria for rivers and streams were used
(Table 7-6; http://www2.epa.gov/nutrient-policv-data/ecoregional-criteria-documents').
Florida has numeric nutrient criteria for TN for most of the state. The compiled state WQC
vary greatly in spatial resolution and N forms addressed. Mississippi applies only one
criterion for the entire state. By contrast, California has a patchwork of criteria designated
by regional water boards.
U.S. EPA Aggregate Level III Omernik Ecoregion Nutrient Criteria for Rivers and Streams
(from U.S. EPA Ecoregional Nutrient Criteria Documents for Rivers and Streams; available
at http://www2.epa.gov/nutrient-policv-data/ecoregional-nutrient-criteria-documents-rivers-
and-streams) were developed for states to provide a more spatially even set of standards for
TN and chlorophyll a (Table 7-7). The regionally based criteria are designed to reflect
characteristics such as soils, vegetation, climate, geology, and land cover, which are
relatively similar within each ecoregion. Pristine or minimally impacted waters from each
region are used as a basis for developing ecoregion-specific nutrient criteria. The aggregate
ecoregion criteria are mapped in Figure 7-4 and Figure 7-5.
The ambient water quality criteria for NH3 was recently updated to reflect the sensitivity of
freshwater unionoid (order Unionoida) mussels to this nutrient (U.S. EPA. 2013a). The
acute criterion is 17 mg total NH3 -N (TAN)/L, and the chronic criterion is 1.9 TAN/L at
pH 7 and temperature 20°C.
7-59
DRAFT: Do Not Cite or Quote

-------
Table 7-6 Numeric nutrient water quality criteria for rivers/streams by state (all
values in mg/L).
State
Subregion
TN
Nitrate
(as N)
Nitrite
(as N)
Nitrate +
Nitrite
(as N)
Ammonia
(NHs)
Chlorophyll a
Alabama
Alaska
-

10
1
10


Arizona
-
1.1-3.0C
10
1



Arkansas
California11
SWRCB Region 1

10




Californiad
SWRCB Region 2

10
1
10


California11
SWRCB Region 3

10




California11
SWRCB Region 4

8
1
10


California11
SWRCB Region 5
0.31




0.0018
California11
SWRCB Region 6
0.38




0.00178
California11
SWRCB Region 7

10

10


California11
SWRCB Region 8
0.38
10



0.00178
Colorado®
Remainder of state

10
0.05

0.02

Colorado®
South Platte Basin


0.5



Connecticut
Delaware
-

10




Florida'
Panhandle West
0.67




0.02
Florida'
Panhandle East
1.03




0.02
Florida'
North Central
1.87




0.02
Florida'
Peninsular
1.54




0.02
Florida'
West Central
1.65




0.02
Florida'
South Florida





0.02
Georgia9
-
4





Hawaii
June 2018
7-60
DRAFT: Do Not Cite or Quote

-------
Table 7-6 (Continued): Numeric nutrient water quality criteria for rivers/streams
by state (all values in mg/L).
State
Subregion
TN
Nitrate
(as N)
Nitrite
(as N)
Nitrate +
Nitrite
(as N)
Ammonia
(NHs)
Chlorophyll a
Idaho
Illinois
-

10




Indiana
-

10
1
10


Iowa
-

10
1



Kansas
-

10

10


Kentucky
-

10
1



Louisiana11
Rest of the state






Louisiana11
Ecoregion 9

0.69



0.00093
Louisiana11
Ecoregion 10

0.76



0.0021
Maine
-

10




Maryland
Massachusetts —
Michigan
-

10




Minnesota'
Central River
Region





0.018
Minnesota'
North River Region





0.007
Minnesota'
South River Region





0.035
Minnesota'
Remainder of state






Mississippi
-

10




Missouri
Montana
-

1

1


Nebraska
-

10
1



Nevada
New
Hampshire
New Jersey
New Mexico
June 2018
7-61
DRAFT: Do Not Cite or Quote

-------
Table 7-6 (Continued): Numeric nutrient water quality criteria for rivers/streams
by state (all values in mg/L).
State
Subregion
TN
Nitrate
(as N)
Nitrite
(as N)
Nitrate +
Nitrite
(as N)
Ammonia
(NHs)
Chlorophyll a
New York
-

10

10


North Carolina -
North Dakota




10


Ohio
Oklahoma'
-





0.01
Oregon
-

10



0.015k
Pennsylvania
-


10



Rhode Island
South Carolina —
South Dakota
Tennessee
-

10




T exas
Utah
-

10




Vermont1
Class A(1) and A(2)
waters above 2,500

0.2




Vermont1
Class AC 1 ¦) and A(2)
waters below 2,500

2




Vermont1
Class B waters

5




Virginia
-

10




Washington"1
Ecoregion 1
0.31




0.0018
Washington"1
Ecoregion 2
0.12




0.00108
Washington"1
Ecoregion 3
0.38




0.00178
West Virginia
-

10




Wisconsin
June 2018
7-62
DRAFT: Do Not Cite or Quote

-------
Table 7-6 (Continued): Numeric nutrient water quality criteria for rivers/streams
by state (all values in mg/L).
State
Subregion
TN
Nitrate
(as N)
Nitrite
(as N)
Nitrate +
Nitrite
(as N)
Ammonia
(NH3) Chlorophyll a
Wyoming
-

10
1
10

L = liter; mg = milligram; N = nitrogen; NH3 = ammonia; TN = total nitrogen.
SWRCB = State Water Resources Control Board
aRegulatory status as of November 2017. Subject to change. In addition to codified criteria, select waterbodies may also have
unique criteria developed as Clean Water Act Section 303(d) Total Maximum Daily Load (TMDL) impairment restoration
planning, See https://www.epa.aov/nutrient-policv-data/state-proaress-toward-developina-numeric-nutrient-water-gualitv-
criteria.
b10 mg/L for Nitrates and 1 mg/L for Nitrites are nonenforced public health U.S. EPA recommendations for Drinking Water
sources.
°Arizona R18-11-109F (varies by waterbody) http://apps.azsos.gov/public services/Title 18/18-11.pdf.
dCalifornia State Water Resources Control Board, https://www.waterboards.ca.gov/water issues/programs/nitrate project/.
eColorado DPEH. https://www.colorado.gov/pacific/cdphe/water-gualitv-control-commission-regulations.
'Florida DEP. https://floridadep.gov/dear/water-gualitv-standards/content/numeric-nutrient-criteria-development.
9Georgia 391-3-6-.03 Water use Classifications and Water Quality Standards
https://epd.georgia.gov/sites/epd.georgia.gov/files/related files/site page/EPA Approved WQS May 1 2015.pdf.
Louisiana Administrative Code (LAC) Title 33. https://www.epa.gov/sites/production/files/2014-12/documents/lawgs.pdf.
'Minnesota Specific Water Quality Standards for Class 2 Waters. Minnesota Administrative Rule 7050.0222. Subpart 2.
https://www.revisor.mn.gov/rules/?id=7050&view=chapter#rule.7050.0222.
'Oklahoma Water Quality Standards_785:45-5-10 (standard for any waterbody designated Sensitive Public and Private Water
Supplies [SWS]) https://www.epa.gov/sites/production/files/2014-12/documents/okwgs chapter45.pdf.
kOregon Water Quality Standards. OAR 340-41-0019 (1)(B)
http://pweb.crohms.org/tmt/wgnew/state standards/or/2004 Chapter 340 Div 041 .pdf.
'Vermont Water Quality Standards Environmental Protection Rule Chapter 29(a).
https://www.epa.gov/sites/production/files/2014-12/documents/vtwgs.pdf.
"Washington. State cites U.S. EPA Nutrient Aggregate Ecoregion Rivers and Streams criteria
https://www.epa.gov/sites/production/files/2016-06/documents/npwdr complete table.pdf.
June 2018
7-63
DRAFT: Do Not Cite or Quote

-------
Table 7-7 U.S. EPA aggregate Level III ecoregion nutrient criteria (all values in
mg/L; U.S. EPA ecoregional nutrient criteria documents for rivers
and streams).
Ecoregion
TN
Chlorophyll a
1
0.31
0.0018
2
0.12
0.00108
3
0.38
0.00179
4
0.56
0.0024
5
0.88
0.003
6
2.18
0.0027
7
0.54
0.0015
8
0.38
0.00063
9
0.69
0.00093
10
0.76
0.0021
11
0.31
0.00161
12
0.9
0.0004
TN = total nitrogen.
June 2018
7-64
DRAFT: Do Not Cite or Quote

-------
L = liter; mg = milligram.
Figure 7-4 Total nitrogen criterion values by ecoregion.
L = liter; mg = milligram.
Figure 7-5 Chlorophyll a criterion values by ecoregion.
June 2018
7-65
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
7.1.7.
Freshwater Biogeochemistry Summary
In the 2008 ISA, the body of evidence was sufficient to infer a causal relationship
between N and S deposition and the alteration of biogeochemical cycling of N and C in
freshwater ecosystems, and between acidifying deposition and changes in
biogeochemistry of fresh waters. The strongest evidence for a causal relationship between
acidifying deposition and aquatic biogeochemistry comes from studies of changes in
surface water chemistry, including concentrations of SO42 , NO3 . inorganic Al, Ca, sum
and surplus of base cations, ANC, and surface water pH. Data from long-term
monitoring, experimental manipulations, and modeling studies provide consistent and
coherent evidence for biogeochemical changes associated with acidifying N and S
deposition. As reviewed in the 2008 ISA and newer studies, biogeochemical processes
and surface water chemistry are influenced by characteristics of the catchment and the
receiving waters. Atmospheric deposition affects the chemistry, and often also the
biology, of each type of freshwater environment. These waters may be chronically
acidified or subject to occasional episodes of decreased pH, decreased ANC, and
increased inorganic Al concentration. Increased N deposition to freshwater systems via
runoff or direct atmospheric deposition, especially to N limited and N and phosphorus (P)
colimited systems, can stimulate primary production. New information on
biogeochemical cycling of N and S, acidifying deposition effects on biogeochemical
processes and changes to chemical indicators of surface water chemistry associated with
acidification and N nutrient enrichment is consistent with the conclusions of the 2008
ISA, and the body of evidence is sufficient to infer a causal relationship between N
and S deposition and the alteration of freshwater biogeochemistry.
Indicators of altered N cycling include changes in the concentrations of NO3 in surface
waters. It was well known at the time of the 2008 ISA that key processes such as
nitrification and denitrification are quantitatively important portions of the N cycle and
that they can be influenced by atmospheric inputs of oxidized and reduced N. More
recent research has further substantiated these earlier findings and provided additional
quantitative context. Some new research suggests that denitrification may, in some
situations, produce more N2O in relationship to surface water NO3 concentration than
was previously widely recognized. The quantity and timing of NO3 leaching into surface
waters is an indicator of terrestrial N cycling in the associated watershed. The
concentration of NO3 in drainage water provides an index of the balance between
removal and addition of N to terrestrial ecosystems. Studies of several types have been
conducted in recent years to elucidate these processes, including experimental studies,
isotopic analyses, monitoring, and observational studies.
June 2018
7-66
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
As stated in the 2008 ISA, decreases in base cation concentrations in surface waters in the
eastern U.S. over the past two to three decades have been ubiquitous and closely tied to
trends in SOr concentrations in surface waters. As reported in the 2008 ISA, lake and
stream ANC values decreased throughout much of the 20th century in a large number of
acid-sensitive lakes and streams throughout the eastern U.S. This effect has been well
documented in monitoring programs, paleolimnological studies, and model simulations.
As stated in the 2008 ISA, the concentration of dissolved inorganic monomeric A1 in
surface waters is an especially useful indicator of the adverse impacts of acidifying
deposition. Since the 2008 ISA, several monitoring studies have reported decreases in
inorganic A1 suggestive of chemical recovery of surface waters lYBaldigo et al.. 2016;
Driscoll et al.. 2016; Strock et al.. 2014; Warbv et al.. 2008); Appendix 7.1.5.11.
Monitoring data have proven to constitute a reliable means with which to confirm
environmental models and evaluate damage/recovery of ecosystems in response to
changes in acidic atmospheric deposition. A number of freshwater monitoring studies
have documented ecosystem damage and recovery caused by acidifying deposition of S
and/or N. Many of these studies have been conducted in the U.S., especially in the
northeast and the Appalachian Mountains. Since 2008, TIME, LTM, and other long-term
studies have documented and quantified the responses of surface waters to changes in
acidic deposition and other ecosystem drivers. Although many of these monitoring
programs were in existence at the time of the 2008 ISA and were considered in that
analysis, more recent publications reflect the longer period of monitoring record and
strengthen previous conclusions.
With long-term decreases in atmospheric S deposition, the effects of future increases in
precipitation that may occur in some areas in response to climate change will likely
become increasingly important in regulating the amount of SOr mobilized from internal
watershed sources. A number of S cycling studies (Rice et al.. 2014; Mitchell et al.. 2013)
have emphasized the importance of S adsorption and desorption and their interactions
with soil pH. In addition, internal watershed sources of S, which were earlier believed to
be relatively minor in the northeastern U.S., have and will likely continue to become
proportionately more important as S deposition continues to decline.
Reductions in SOx deposition have not consistently resulted in increases of ANC in
surface water. The quantitatively most important component of the overall surface water
acidification and chemical recovery responses has been change in base cation supply.
This was highlighted in the assessment of Charles and Christie (1991) and in the 2008
ISA. Base cations are added to watershed soils by weathering of minerals and
atmospheric deposition, and are removed by uptake into growing vegetation or by
leaching. Acidic deposition increased leaching of base cations, as SOr anions in soil
7-67
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
solution carried along base cations to maintain the charge balance. In watersheds that
have received high levels of historical acidic deposition, current exchangeable
concentrations of Ca2+ and other base cations are substantially reduced from likely
preindustrial levels, having been depleted by many years of acidic deposition. This base
cation depletion constrains ANC and pH recovery of surface waters, as described in the
2008 ISA. Recent results have further corroborated earlier findings and have included
experiments, modeling, and gradient studies.
Recent research has described an ecosystem recovery response to decreasing SOr
deposition that was not a major focus of the 2008 ISA: DOC increases in surface water.
Such changes have been attributed to increased atmospheric CO2 concentration, climate
warming, decreased S deposition with associated changes in water pH and ionic strength,
and hydrologic changes associated with drought and precipitation. Although not all
increases in DOC are directly caused by decreases in acidifying deposition, any changes
in DOC concentration or properties can impact the acid-base chemistry of surface waters
and perhaps the composition of aquatic biota. It has been recognized that surface water
DOC concentrations had decreased to some extent as a result of acidification, and that
DOC would likely increase with recovery. However, the strength of this response and the
magnitude of DOC changes have exceeded scientific predictions. Recent research on this
topic has been diverse and has included experiments, observation, isotope studies, and
synthesis and integration work. Overall, these studies illustrate large increases in DOC
with acidification recovery in some aquatic systems. Increases in DOC constrain the
extent of ANC and pH recovery, but decrease the toxicity of dissolved A1 by converting
some of it from inorganic to organic forms (Lawrence et al.. 2013). However, DOC is not
an indicator of recovery everywhere; some recovering sites have not shown increasing
trends in DOC.
Taken together, results of recent acid-base chemistry studies in the northeastern U.S.
confirmed the previously observed pattern of gradual surface water ANC and pH
recovery, in some cases more marked decrease in inorganic Al concentrations, and
important interactions with DOC. It is important to recognize that recent changes in
atmospheric S and N deposition, although substantial in many areas of the U.S., have not
occurred in a vacuum. Other important ecosystem drivers have also changed, including
various kinds of disturbance, and these changes have varied by region. The potential
importance of such factors was known at the time of preparation of the 2008 ISA. More
recent work has further confirmed the importance of these linkages, especially those
related to climate (Appendix 13).
7-68
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
7.2. Biogeochemistry of Nitrogen in Estuarine and Near-Coastal
Systems
In the 2008 ISA, the evidence was sufficient to infer a causal relationship between N
deposition and alterations in the biogeochemical cycling of N and carbon (C) in estuarine
and near-coastal marine systems. Since the 2008 ISA, additional studies have quantified
atmospheric N deposition to estuaries, especially along the Atlantic coast. New evidence
shows that a trend from dominantly oxidized N inputs to dominantly reduced forms of N
input have ramifications for N cycling and biological response in some estuaries. There is
additional evidence that N alters biogeochemistry of coastal ecosystems, especially in
regard to microbial-mediated N transformations, which play large roles in N within
estuaries. Research and modeling since the 2008 ISA have shown that many of these N
processes such as dissimilatory NOs reduction to NH4+ (DNRA) are more important in
the estuarine environment than previously thought, and that rates of N cycling can be
highly variable. Key processes affected by N loading include nitrification and
denitrification. Monitoring of coastal areas shows that excess nutrient inputs continues to
be a widespread problem in many parts of the U.S. The body of evidence is sufficient to
infer a causal relationship between N deposition and the alteration of
biogeochemistry in estuarine and near-coastal marine systems, which is consistent
with the findings of the 2008 ISA.
Recently, elevated nutrient loading to coastal areas has been hypothesized to contribute to
coastal acidification. Ocean acidification is already occurring from dissolution of rising
atmospheric CO2. Nitrogen enrichment is expected to worsen this acidification when
algal biomass and excess organic matter from increased primary production and
allocthonous inputs decomposes, and the CO2 produced by decomposition undergoes
dissociation and production of hydrogen ions, making the water more acidic
(Appendix 7.2.4V Respiration of living algae and seagrasses during the night can also
contribute to acidification. Nutrient-enhanced coastal acidification was not discussed in
the 2008 ISA. The body of evidence is sufficient to infer a likely causal relationship
between N deposition and increased nutrient-enhanced coastal acidification.
Atmospherically deposited N to watersheds, along with other nonatmospheric sources of
N, influence processes that operate along the headwater to ocean continuum. Total N
loading to estuaries includes riverine transport of N and direct deposition of N to the
estuary itself. This influx of N alters N and C cycling and leads to estuary eutrophication,
the process of nutrient over-enrichment. Eutrophic systems are characterized by an
increase in the rate of supply of organic matter (primary production and organic C
accumulation) in excess of what an ecosystem is normally adapted to processing (Diaz et
al.. 2013; Nixon. 1995). Estuary eutrophication is indicated by water quality
June 2018
7-69
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
deterioration, including development of hypoxic zones, species mortality, and formation
of harmful algal blooms (HABs). Biological indicators of estuarine condition
(e.g., chlorophyll a, HABs, macroalgae, submerged aquatic vegetation [SAV]) are
described in Appendix 10. Andersen et al. (2006) suggested eutrophication be defined as
"the enrichment of water by nutrients, especially N and P and organic matter, causing
increased growth of algae and higher forms of plant life to produce an unacceptable
deviation in structure function and stability of organisms present in the water and to the
quality of water concerned, compared to reference conditions." This appendix considers
biogeochemical processes affected by N loading, a topic not covered in detail in the 2008
ISA. Although this ISA includes both N and S deposition, seawater contains high
concentrations of SOr so atmospheric inputs of S are unlikely to contribute substantially
to biogeochemistry in coastal areas and will not be discussed further in the following
sections.
Estuarine biogeochemistry is complicated because it directly controls more than just the
N cycle; the response to N loading resulting in eutrophication impacts the chemical
cycling of metals and DO (Appendix 7.2.3). redox conditions, pH (Appendix 7.2.4) and
ultimately energy transfer (e.g., food webs from microbes to humans). The response to N
loading is also tightly controlled by the availability of organic matter (C) and its lability
and reactivity. Excess nutrient inputs are occurring within the context of other stressors
such as climate change (Appendix 7.2.6.12) and rising atmospheric CO2, which further
modify coastal biogeochemistry (Donev. 2010). In the complex environment of the
freshwater-to-ocean continuum, there are many chemical and biological indicators of
eutrophic condition. One approach is to measure total nutrient loading and
concentrations; however, these data need to be interpreted in the context of the physical
and hydrological characteristics that determine ecosystem response. Water quality
measures such as pH and DO, along with key biological indicators such as chlorophyll a,
phytoplankton abundance, harmful algal blooms (HABs), macroalgal abundance, and
SAV, can all be used to assess responses to nutrient loading (Table 7-8). The following
sections highlight post-2007 research focusing on N biogeochemistry in estuaries and
near-coastal areas.
Table 7-8 Summary of key indicators of nitrogen enrichment and
nutrient-enhanced coastal acidification.
Endpoint
N-Driven Nutrient
Enrichment
Nutrient-Enhanced
Coastal
Acidification
Section of ISA That
Discusses Endpoint
Chemical indicator
June 2018
7-70
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Endpoint
N-Driven Nutrient
Enrichment
Nutrient-Enhanced
Coastal
Acidification
Section of ISA That
Discusses Endpoint
Dissolved oxygen
X
X
7.2.3. 10.2.4
Water pH
X
X
7.2.4. 10.5
Biological indicator
Chlorophyll a
X

10.2.1
Harmful algal blooms
X

10.2.2
Macroalgal abundance
X

10.2.3
Submerged aquatic vegetation
X

10.2.5

7.2.1. Nitrogen Sources
In the 2008 ISA, it was well understood that the importance of atmospheric deposition as
a cause of estuarine eutrophication is determined by the relative contribution of the
atmospheric versus nonatmospheric sources of N input (U.S. EPA. 2008a').
Atmospherically deposited N from combustion of fossil fuels (NOy), agricultural sources
(NH3, NH4+) and organic N combine with N from other anthropogenic and natural
sources to contribute varying total N inputs in coastal regions. It is the cumulative effects
from multiple sources that contributes to ecosystem enrichment (Paerl et al.. 2002).
Atmospheric deposition of reduced N has increased relative to oxidized N in parts of the
U.S. including the East and Midwest in the last few decades (Appendix 2). and this trend
is expected to continue in the future under existing emissions controls (Pinder et al..
2008; U.S. EPA. 2008a'). Thus, the form of inorganic N input to coastal areas is changing
over time. This has ramifications for N cycling. The increase in highly bioreactive
reduced N from deposition and other sources is often a preferred form of N for
phytoplankton including unwanted or harmful species (Appendix 10.3.3'). Deposition of
oxidized and reduced forms of N are detailed in Appendix 2.
In many places throughout the U.S., nonpoint sources are now the dominant sources of N
to water bodies (Howarth. 2008a. b; Howarth et al.. 2002). Releases of N from
agricultural, urban, and mixed land uses comprise a significant portion of the nonpoint
sources in many watersheds (Birch etal.. 2011; Alexander et al.. 2008). The Chesapeake
Bay is an example of a well-studied coastal system where N fluxes have been relatively
well characterized (atmospheric inputs represent approximately 25% of total N budget;
Figure 7-6). In management of coastal eutrophication, both point and nonpoint sources
have been identified as targets for control of N inputs (Stephenson et al.. 2010; Paerl et
al.. 2002). Nonpoint sources include diffuse agricultural runoff and wet and dry
June 2018
7-71
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
deposition from the atmosphere. Increased respiration caused by N enrichment may
exacerbate coastal ocean acidification through alteration of the C cycle. Nitrogen sources
also cause production of excess organic matter that feed respiration by microbial
decomposers. In estuaries adjacent to areas of coastal upwelling, such as some locations
along the Pacific coast of the U.S., oceanic inputs of nutrients can also represent an
important source of N to estuarine waters (Brown and Ozretich. 2009).
Estimates of the contribution of atmospheric deposition to the overall estuary N load
exhibit wide variability due in part to the difficulty associated with quantifying multiple
agricultural and mobile and stationary fuel combustion emissions sources and N
transformations and losses as the N moves from the terrestrial watershed to and within
the estuary (U.S. EPA. 2008a; NRC. 2000). Despite high variability, it appears that
atmospheric sources of N loading to estuaries can be quantitatively important, based on
studies reviewed in the 2008 ISA and newer studies (Table 7-9). For example,
atmospheric N loads to estuaries in the U.S. were estimated to be as high as 72% for St.
Catherine's-Sapelo estuary, Georgia (Castro et al.. 2003). In another study considering
atmospheric deposition of N to the East Coast and Gulf of Mexico, direct deposition of N
was estimated to contribute from 10 to over 40% of N loading to estuarine, coastal, and
open marine waters (Paerl et al.. 2002). In Chesapeake Bay, studies reviewed in the 2008
ISA indicated that atmospheric deposition makes a substantial contribution (up to 50%)
to the overall N budget (Howarth. 2008b; Bover et al.. 2002).
7-72
DRAFT: Do Not Cite or Quote

-------
Chemical Nitrogen Cascade: Chesapeake Bay Watershed
NOx Emissions

Utilities
52,000
Industry
43,000
Mobile Sources
170,000
Other Sources
14,000
\ H, Emissions
Non-Agriculture 22,000
\ Additions to Land
Agriculture 370,000
Urtiatt and M ixed Open
I-and Uses 62,000
N Additions to Water
Point Sources 26.000
Arrow width indicates size of flux.
Arrow color indicates origin of
nitrogen flux
*	Red - Air
*	Green - Land
*	Blue - Water
All estimates are in metric tonnes N
per year
N.E, = No Estimate Available
¦
¦
¦
Delivered to

Delivered to

[delivered to
Bay

Bav

Bav
Source: Birch et al. (2011)
Figure 7-6
Chemical nitrogen cascade in the Chesapeake Bay watershed
(metric tons/year).
June 2018
7-73
DRAFT: Do Not Cite or Quote

-------
Table 7-9 Summary of studies quantifying atmospheric nitrogen contribution
to total nitrogen in coastal areas via watersheds and/or direct
deposition to estuary surface waters.
Region
Total N Loading Due to Atmospheric Deposition
Model
Reference
U.S. coastal areas
N deposition to land surfaces that is subsequently
exported from watersheds to the coastal zone is
responsible for 17 to 21 % of the total N input.
NEWS
SPARROW
McCrackin et al.
(2013)
Gulf of Mexico in the Atmospheric deposition contribution, which may	SPARROW using Robertson and
Mississippi/Atchafalaya include volatilized losses from natural, urban, and	CMAQ total	Saad (2013)
River basin	agricultural sources, was estimated to be 26% of	deposition
total N transported to the Gulf of Mexico.
Northeast and	Identified wet deposition to the watershed as the SPARROW Moore et al.
mid-Atlantic coastal dominant source of N to the estuaries of the	(2011)
region	Connecticut, Kennebec, and Penobscot rivers, but
the third largest source (20%) for the region as a
whole, after agriculture (37%) and sewage and
population-related sources (28%).
Narragansett Bay
Combined direct atmospheric deposition to the
estuary and atmospheric deposition to the
watershed were responsible for 20% of N loading to
the bay.
Vadeboncoeur et
al. (2010)
Waquoit Bay Estuaries
in Cape Cod
Since 1990 wastewater N loads have increased
about 80% while loads from atmospheric deposition
have decreased by about 41% with no change in
total loading on a decadal scale.
NLM
Valiela et al.
(2016)
Small-to-medium sized
estuaries of southern
New England
Direct atmospheric deposition to estuary surface
averaged 37%, and indirect atmospheric deposition
via the watershed averaged 16% of total N loading,
although the percentage varied widely for each
individual estuary.
NLM
Latimer and
Charpentier
(2010)
Tampa Bay
Direct and indirect atmospheric loading were
estimated to be 30 and 41%, respectively, of total N
loading to the bay.
WDT
Poor et al.
(2013b)
Direct and indirect atmospheric loading were
estimated to be 17 and 40%, respectively, of total N
loading to the bay.
CMAQ modified Poor et al.
with University of (2013a)
California Davis
aerosol module
Chesapeake Bay
Atmospheric loading is 24% of total N loading.
Birch et al.
(2011)
Half of the atmospheric source of N to the watershed Chesapeake
Linker et al.
originates outside of the watershed; indirect loading
of atmospheric N via watershed inputs are larger
than direct loading to tidal waters.
Airshed Model, (2013)
combining CMAQ
and regression
model for wet
deposition
June 2018
7-74
DRAFT: Do Not Cite or Quote

-------
Table 7-9 (Continued): Summary of studies quantifying atmospheric nitrogen
contribution to total nitrogen in coastal areas via
watersheds and/or direct deposition to estuary surface
waters.
Region
Total N Loading Due to Atmospheric Deposition
Model
Reference
Gulf of Mexico in the
Mississippi/Atchafalaya
River basin
Atmospheric deposition to watersheds in the basin
contributed about 16% of the total N load, second to
N from corn and soybean production (52%).
SPARROW
Alexander et al.
(2008)
Pacific Northwest,
Yaquina Bay
Direct deposition represented only 0.03% of N inputs
and watershed inputs of N fixing red alder (Alnus
rubra) was a greater source of N to the watershed
than atmospheric deposition (8%).

Brown and
Ozretich (2009)
Key pre-2008 literature
34 Atlantic and Gulf
coast estuaries
The contribution of atmospheric deposition
(including directly onto the water surface and onto
the watershed) was 7-72% of the total N.
WATERS-N
Castro et al.
(2001);Castro et
al. (2003)
10 estuaries along the
U.S. east coast
Total atmospheric inputs (watershed runoff plus
direct deposition to the surface of the estuary)
accounted for 15 to 42% of total N inputs. In four of
the estuaries direct deposition was 35 to 50% of the
total atmospheric N inputs.

Castro and
Driscoll (2002)
Chesapeake Bay
Atmospheric deposition makes a substantial
contribution (about 25%) to the overall N budget.

Howarth (2007)
16 northeastern river
basins
Atmospheric deposition averaged 31% of total N
inputs; values for watersheds in northern New
England were substantially higher and atmospheric
deposition dominated. Estimated riverine export of N
amounted to 25% of total N inputs.

Bover et al.
(2002)
Based on 34 Atlantic
and Gulf coast
estuaries
In systems with watershed area to estuarine surface
area ratios greater than 15, N deposition is not as
important (comprising less than 25% total N input)
as in estuaries that are large relative to their
watershed.

Castro et al.
(2001)
CMAQ = Community Multiscale Air Quality; N = nitrogen; NEWS = Nutrient Export from Watersheds; NLM = Nitrogen Loading
model; SPARROW = Spatially Referenced Regressions on Watershed Attributes; WATERS-N = Watershed Assessment Tool for
Evaluating Reduction Scenarios for Nitrogen; WDT = Water Deposition Tool.
June 2018
7-75
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
In a series of studies reviewed in the 2008 ISA, the model Watershed Assessment Tool
for Evaluating Reduction Scenarios for Nitrogen (WATERS-N) estimated that the
contribution of atmospheric deposition (including directly onto the water surface and
indirectly onto the watershed) was 7-60% of the total N input to 34 Atlantic and Gulf
coast estuaries (Castro et al.. 2003; Castro et al.. 2001). Nitrogen deposition contributed
20-86% (average = 48%) of the total N input to 22 watersheds having watershed area to
estuarine surface area ratio less than 15. In systems with a ratio greater than 15, N
deposition is not as important, comprising less than 25% of total N input across
12 systems (Castro et al.. 2001).
Since the 2008 ISA, additional modeling studies have estimated the amount and
proportion of current and future N loading expected to result from atmospheric deposition
(Table 7-1). The Nutrient Export from Watersheds (NEWS) and SPARROW models
agreed that 62-81% of N delivered to the coastal zone in the continental U.S. is
anthropogenic in source (McCrackin et al.. 2013). Model results showed that atmospheric
N deposition that is subsequently transported to estuaries represents 17-21% of the total
N exported to the coastal zone (McCrackin et al.. 2013; Moore etal.. 2011). Using
SPARROW atmospheric N deposition was identified as the dominant source of delivered
N to estuaries along the Atlantic Coast and the Gulf of Mexico (McCrackin et al.. 2013).
Using SPARROW, the atmospheric deposition contribution (which may include
volatilized losses from natural, urban, and agricultural sources) was estimated to be 26%
of total N transported to the Gulf of Mexico in the Mississippi/Atchafalaya River basin
(Robertson and Saad. 2013). Moore et al. (2011) found that SPARROW identified
atmospheric deposition to watersheds as the dominant source of N to the estuaries of the
Connecticut, Kennebec, and Penobscot rivers, but the third largest source (20%) for the
Northeast and mid-Atlantic coastal region as a whole, after agriculture (37%) and sewage
and population-related sources (28%). In contrast, atmospheric deposition is a less
substantial source of N to some estuaries, especially in the Pacific Northwest. In Yaquina
Bay estuary, OR, direct deposition represented only 0.03% of N inputs and watershed
inputs of N fixing red alder (Alnus rubra) trees was a larger (8%) source of N to the
watershed than atmospheric deposition (Brown and Ozretich. 2009). In Yaquina Bay
estuary, the ocean is the primary source of N during the dry season and the river is the
primary source during the wet season.
A modeling study conducted by Vadeboncoeur et al. (2010) estimated that the third
largest source of N loading (via the watershed and directly to the water body) to
Narragansett Bay in the year 2000 was atmospheric deposition (20%). Future scenarios
suggested that very aggressive reductions in both fertilizer use and atmospheric
deposition would be needed for Narragansett Bay to return to early twentieth century
levels of N loading (Vadeboncoeur et al.. 2010). Latimer and Charpentier (2010) applied
7-76
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
the Watershed N Loading Model (NLM) to small- to medium-sized estuaries of southern
New England. Direct atmospheric deposition to the water surface made up an average of
37% of the N input, although the percentage varied widely for individual estuaries
(Latimer and Charpentier. 2010). In the same study, wastewater represented 36% of the
total N inputs. Indirect atmospheric deposition within the watershed was estimated to be
16% and the fertilizer component was 12%. The relative magnitudes of these source
types varied across estuaries. In another modeling study with NLM using data going back
to 1990, atmospheric deposition decreased by about 41% while wastewater inputs have
increased 80% with a net result that total loads have not changed on a decadal scale in the
Waquoit Bay estuaries in Cape Cod, MA (Valielaetal.. 2016). Using the Watershed
Deposition Tool (WDT), Poor et al. (2013b) estimated direct atmospheric loading of
1,080 metric tons of N and 1,490 metric tons of N from indirect atmospheric loading
(assuming a watershed-to-bay transfer rate of 18%) for Tampa Bay using data from 2002.
Thus, atmospheric sources were an estimated 71% of the total N loading. Key studies
from the 2008 ISA on atmospheric N loading to estuaries are summarized in Table 7-9.
along with newer estimates.
Near NH3 emission sources, such as large mammal and poultry livestock operations, N
deposition is expected to increase 10-40% by 2020 according to different regulation and
reduction scenarios (Pinder et al.. 2008). Many of the agricultural emission sources are
located in sensitive watersheds such as the Chesapeake Bay and Pamlico Sound, NC
(Pinder et al.. 2008). Birch et al. (2011) estimated that 24% of the reactive N that reached
Chesapeake Bay originated as emissions to the atmosphere that were then deposited in
the watershed to land and water, 57% came from terrestrial N additions from nonpoint
source runoff, and the remaining 19% was estimated to come from direct discharges of N
to freshwater ecosystems. The results of a modeling study that integrated the airshed,
watershed, and estuary indicated that atmospheric deposition represented one of the
largest inputs of N to the Chesapeake Bay watershed, and about half of it originated
outside the watershed (Linker et al.. 2013). Atmospheric N loading to tidal waters was
included in the Chesapeake Bay 2010 TMDL, the first time that atmospheric deposition
has been included in a plan for nutrient load reduction (U.S. EPA. 2010). The explicit
TMDL N allocation was determined to be 7.1 million kg per year of total N atmospheric
deposition loads directly to Chesapeake Bay and tidal tributary surface waters. It was
based on the 2020 CMAQ N deposition scenario described by Linker etal. (2013).
Appendix 2 provides more detailed discussion of N source apportionment.
Although the geographic scope of this ISA is limited to estuaries and near-coastal areas
of the U.S., some evidence indicates that the open ocean may be affected by atmospheric
deposition of N (Ren et al.. 2017; Kim etal.. 2011). Ren et al. (2017) provided 15N stable
isotope evidence that N emissions have changed the biogeochemical cycling of N in the
7-77
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
open ocean in nutrient-limited low-latitude regions such as the northwestern Pacific
Ocean. Changes in 15N overtime were documented in coral skeletons over a period of
45 years at Dongsha Atoll in the northern South China Sea. The proximity of this water
body to anthropogenic N emissions sources in Asia and its depth (1,300 m) both
contribute to its sensitivity to N deposition and changes in nutrient cycling.
7.2.2. The Estuary Environment
Atmospherically deposited N, along with other sources of N to coastal systems,
influences uptake and conversion processes that operate along the freshwater-to-ocean
continuum KPacrl and Piehler. 2008; Seitzinger et al.. 2006); Figure 7-71. Conditions in
upstream terrestrial systems, such as nutrient loading or land use, can exert strong
influences on coastal habitats and processes (Ruttcnbcrg and Granek. 2011) Appendix 4.
In addition to inputs ofN from outside sources, the varying rates of different N cycling
processes, as well as C availability and reactivity within estuaries themselves, can also
affect the magnitude of eutrophication experienced as a result of external N enrichment
(Newell et al.. 2016; Anderson et al.. 2014b; Smyth et al.. 2013; Crowe et al.. 2012).
Estuaries are heterogeneous environments characterized by physicochemical gradients of
salinity and both naturally and anthropogenically derived nutrients. Water quality in
estuaries is highly variable due to physical and chemical aspects of each estuary
(Appendix 10.1.4) and factors such as temperature and precipitation (Rheuban et al..
2016). At the upstream end of an estuary, the water is primarily fresh much of the time.
Nonpoint source runoff of N from the land surface, only a part of which is of atmospheric
origin (mainly deposition to land and subsequently leached to the river water), dominates
new N inputs. Downstream, freshwater inflows gradually mix with salt water to form
mesohaline segments of the estuary. These changes in salinity impact the ionic strength
of the water. Further along the salinity gradient, a significant fraction of the terrestrial
biologically available N load is assimilated by phytoplankton and benthic flora, removed
by microbes in the process of denitrification or advected to the ocean (Pacrl et al.. 2002).
Direct atmospheric N inputs also impact the lower estuary, although the relative
importance of N deposition in this zone is uncertain and variable (Pacrl et al.. 2002). In
many estuary and sound ecosystems, primary production and phytoplankton biomass are
highest at mid-estuary locations, where N loads and decreasing rates of flushing
(increasing residence times) overlap.
June 2018
7-78
DRAFT: Do Not Cite or Quote

-------
Atmospheric
Urban/
Agriculture/ *
Industrial
Runoff^ (Aquaeulwre
Coastal
Ocean
Open
Ocean
Estuary/ Son nil
rroundwater
i Hypoxia

Chi a = chlorophyll a; DNF = denitrified; N = nitrogen; N2 = nitrogen; NF = nitrogen (N2) fixation.
Source: From Paerl and Piehler (2008).
Figure 7-7 Schematic diagram illustrating sources, transformations, and fate
of nitrogen along the estuary-to-ocean continuum. Surface,
subsurface, and atmospheric pathways of externally supplied or
new nitrogen inputs attributable to anthropogenic activities are
shown as internal nitrogen cycling. The combined anthropogenic
nitrogen inputs are shown as a thick arrow (upstream), which
decreases in thickness downstream as a portion of the nitrogen
inputs settles to the bottom sediments and is buried and/or
denitrified. Nitrogen (N2) fixation is a biologically mediated new
nitrogen input. The linkage of anthropogenically enhanced
nitrogen inputs to accelerated primary production or
eutrophication and its trophic and biogeochemical fate are also
shown.
7.2.3.	Dissolved Oxygen and Hypoxia
1	Low oxygen (hypoxia) or the absence of oxygen (anoxia) in coastal waters have
2	implications for N cycling. Oxygen depletion mainly occurs in bottom waters under
3	stratified conditions. The extent of hypoxia in U.S. waters and effects of low DO on biota
4	are discussed in Appendix 10.2.4. Biogeochemical processes that occur in the sediment
5	of estuarine and near-coastal ecosystems are important to the cycling of N that is
6	deposited or transported to estuarine ecosystems. Inputs of C and a stratified water
7	column that separates well-oxygenated surface water from bottom water and sediments is
8	generally required for formation of hypoxic zones (Jewett et al.. 2010).
9	Eutrophication-induced hypoxia, which has been documented globally, can be
June 2018
7-79
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
characterized by both the duration of the hypoxic event and the ecosystem response (Diaz
et al.. 2013; Diaz and Rosenberg. 2008). Summer hypoxia is most common, followed by
systems that experience periodic O2 depletion that may occur more often than seasonally.
Development of seasonal hypoxia is common in shallow coastal regions that receive high
inputs of nutrients, including N, from coastal rivers. Development of hypoxia is
increasingly becoming a concern throughout the U.S. and internationally (Jcwett et al..
2010; Diaz and Rosenberg. 2008). Middelburg and Levin (2009) provided a review of
coastal hypoxia and linkages with biogeochemistry. Changes in bottom water O2
concentrations influence chemical exchange between sediment and bottom water.
Rabalais et al. (2010) provided a synthesis of knowledge regarding hypoxia in coastal
waters. Hypoxic water masses (less than approximately 30% saturation) are more likely
to occur in marine waters where and when water residence time is long, water exchange
is limited, the water column stratifies, and the production and movement of C to bottom
waters are relatively high. Howarth et al. (2011) reviewed the complex interactions
between biogeochemical cycling, eutrophication, and hypoxia in coastal marine systems.
Biochemical feedbacks under eutrophic conditions accelerate further eutrophication and
hypoxia. In locations where coastal upwelling can be a large source of nutrient loads such
as the Pacific Northwest, advection of upwelled water can introduce hypoxic water into
estuaries that is not or poorly related to anthropogenic eutrophication (Brown and Power.
2011; Brown and Ozretich. 2009).
7.2.4. Estuarine and Near-Coastal pH
The pH of estuarine waters can affect N cycling processes and reflect coastal
acidification. Since the 2008 ISA, a number of papers have identified links between
nutrient enrichment and coastal acidification, and several mechanisms have been
identified. One of the initial studies found that CO2 production during decomposition of
organic matter delivered to coastal zones from rivers experiencing eutrophication has
enhanced the acidification of coastal subsurface waters in the Gulf of Mexico and the
East China Sea (Cai etal.. 2011c). Coupled with dissolution of atmospheric
anthropogenic CO2 into the ocean which has led to long-term decreases in pH, CO2
produced from decomposition of excess organic matter associated with eutrophication
can also lower pH. Production of CO2 by living algae and seagrasses during the night can
also drive acidification. In a eutrophic seagrass ecosystem in Cape Cod, MA, there was a
very pronounced diel pattern of pH, with moderate pH during the day, but by dawn very
acidic waters resulted from overnight respiration of living algae and seagrasses (Howarth
et al.. 2014). During the daytime, CO2 was drawn down through primary production, and
as a consequence the pH steadily climbed. The net effect was to increase acidification,
June 2018
7-80
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
due to slow exchange of CO2 with the atmosphere. This additional CO2 further acidifies
marine waters as it dissociates into carbonate ions and hydrogen ions (Sunda and Cai.
2012; Cai etal.. 2011c; Howarth et al.. 2011). An important additional source of organic
matter leading to overall declines in pH is potentially allochthonous organic matter inputs
that have been increasing in many coastal watersheds from changing land use (Wilson et
al.. 2016; Wetz and Yoskowitz. 2013). Nitrogen-driven eutrophication and
anthropogenically enhanced allochthonous organic matter loading operate
simultaneously.
Biogeochemical modeling of eutrophically driven ocean acidification coupled to data
collected from the Gulf of Mexico, the Baltic Sea, and the East China Sea predicted that
eutrophication would cause a 0.24 to 1.1 unit decrease in pH of bottom waters, and also
that increasing atmospheric CO2 will synergistically amplify eutrophically driven
acidification (Sunda and Cai. 2012; Cai et al.. 2011c'). The increase in atmospheric CO2
dissolution is also expected to alter the N cycle in the ocean, with implications for the
entire ecosystem. Acidification will result in decreased rates of nitrification, which,
combined with expected increasing N deposition, will likely cause the NH44"
concentration of the water to rise (Lefebvre et al.. 2012). Models show that while the
impact of each acidification pathway (N enrichment and atmospheric CO2 dissolution)
may be moderate, the combined effect of the two may be much larger than would be
expected by just adding the effects of each pathway together (Sunda and Cai. 2012; Cai et
al.. 201 lc).
More recent work has identified some similar dynamics in Chesapeake Bay, where
anoxia related to eutrophication is driving acidification and carbonate mineral dissolution
(Cai et al.. 2017b). Ocean acidification has detrimental effects on marine calcifiers and is
projected to alter marine habitat and food webs affecting a wide range of marine
ecosystem processes [(Marshall et al.. 2017; Mostofa et al.. 2016; Andersson et al.. 2015;
Sunda and Cai. 2012; Donev et al.. 2009) Appendix 10.51.
7.2.5. Nitrogen in Surface Waters
In most estuaries, N inputs from atmospheric deposition, wastewater and agricultural
runoff control eutrophication (Valiela et al.. 2016; Howarth et al.. 1996a; Vitousek and
Howarth. 1991) and can be linked to changes in biological indicators of nutrient
enrichment known to be sensitive to shifts in N loading (Appendix 10). In many
estuaries, this occurs with a simultaneous increase in P as N, and P runoff loads are often
coupled, especially where agricultural or human waste sources dominate. In general,
estuaries tend to be N limited (Howarth et al.. 2011; Paerl and Piehler. 2008; Elser et al..
June 2018
7-81
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
2007; Howarth and Marino. 2006; NRC. 2000; Nixon. 1995; Howarth. 1988). However,
more recent studies show that some estuaries are P limited, or colimited by N and P, or
switch seasonally between N and P limitation (Howarth et al.. 2011; Paerl and Piehler.
2008; Howarth and Marino. 2006). To date, several states have developed site-specific or
state-wide numeric nutrient criteria to address nutrient pollution problems
(Appendix 7.2.9). Adopted criteria include TN, NOs as N, DIN as N, or TMDLs. In
some estuaries, especially in the Pacific Northwest, nutrients from upwelling and oceanic
exchange caused by regional wind patterns likely control primary production rather than
anthropogenic nutrient loading (Brown and Ozretich. 2009; Hickev and Banas. 2003).
Quantification of N in the estuarine environment can be highly variable both spatially and
temporally due to the physiochemical gradients within estuaries (Appendix 10.1.4). the
influence of natural and anthropogenic stressors, and factors such as precipitation and
warming (Appendix 10.1.4.1). For example, analysis of 22 years of water quality data
from Buzzards Bay, MA by Rheuban et al. (2016) indicated that climate related stressors
of warming and precipitation appeared to influence ecosystem response to TN over time.
Since the 2008 ISA, additional thresholds of response to N have been identified for
biological indicators (Appendix 10.6).
7.2.6. Nitrogen Cycling
Some of the key processes involved in N cycling in estuarine and near coastal
environments include N sinks such as denitrification and N2 production via anaerobic
ammonium oxidation (anammox), bioavailable N sources such as N fixation and DNRA,
and other N cycling processes including nitrification (Figure 7-8). A major control on
these processes is C availability and reactivity (Plummer et al.. 2015). Benthic DO
concentrations (hypoxia) and S cycling may also affect the occurrence and rates of many
of these N cycling processes. Recent research has shown that many of these N cycling
processes are more important in the estuarine environment than previously thought, and
that the rates of different N cycling processes can be highly variable, depending on
environmental factors such as N availability, organic C source, temperature, seasonality,
oxygen levels, microbial community structure, aquatic habitat type, and others (Newell et
al.. 2016; Plummer et al.. 2015). Some processes were relatively well understood at the
time of the 2008 ISA while the role of others such as DNRA has been characterized only
recently as an important pathway in which N is conserved in coastal ecosystems.
June 2018
7-82
DRAFT: Do Not Cite or Quote

-------
N03- Org-N
Assimilatory
NO3- reduction
(>13)
Water column
Sediment
Derrtnfication (lis
Org-N
\ 		
Assimilation
V
Ammonification {>16)
/ I
N°r~
/ \Nitrification
N2
NO
An amiTK
//
NH4+-
^ /
\ Oissimilatory NO3- reduction
\ / (0.0005)
Notes: numbers in parentheses are rates of the reactions in |jmol/rn2/h as measured by Crowe et al. (2012).
Figure 7-8 New complexities in nitrogen cycling have been detailed since the
2008 ISA as shown in this illustration of the sedimentary N cycle
in the Lower St, Lawrence estuary.
7.2.6.1. Nitrogen Fixation
1	Nitrogen fixation is the microbially mediated conversion of atmospheric nitrogen gas
2	(N2) to a bioavailable form of N, (i.e., ammonia [NH3]). The community of N fixing
3	microorganisms is more diverse in estuarine and coastal waters than previously thought,
4	and N fixation occurs more widely than previously assumed. Advances in research
5	methods for measuring heterotrophic N fixation have led to the discover} that rates of N
6	fixation 111 coastal sediments can cause the sediments to be a significant source of N to
7	the ecosystem (Newell et al.. 2016). Newell et al. (2016) reported on several studies of N
8	fixation rates in coastal sediments (measured as sediment N2 uptake) with rates ranging
9	from 12 to -250 umol N/m2/h. In some cases, these higher rates of N fixation can
June 20 f 8
7-83
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
potentially exacerbate the eutrophic conditions already experienced in an estuary, as was
the case in Narragansett Bay, RI, where up to 30% of the N added to the system came
from N fixation in some years (Newell et al.. 2016; Fulweiler and Heiss. 2014).
Research in the past decade has also sought to determine what environmental factors may
cause sediments to switch from acting as a sink to acting as a net source of N due to
increased N fixation. Organic matter enrichment appears to be one factor. Fulweiler et al.
(2008) noted a "threshold" of organic enrichment at 0.3 g C/m2/day, below which N
fixation took place at high rates but above which the rates of N fixation declined and the
sediments became a net sink of N due to denitrification (Fulweiler et al.. 2008).
In Copano Bay, a shallow bay on the Gulf of Mexico in Texas, periods of drought
coincided with higher rates of N fixation (Bruesewitz et al.. 2013). Thus, base flow
conditions allowed primary production to continue, even when external sources of N via
riverine transport and runoff were very low.
7.2.6.2.	Role of Dissolved Organic Nitrogen
Recent research has shown the importance of measuring DON fluxes, as they can be a
more significant internal source of N from the sediments than previously assumed.
Terrestrial inputs of DON are poorly characterized, and these fluxes are not routinely
included in estuarine nutrient budgets. Alkhatib et al. (2013) found that DON fluxes out
of the sediments, at rates of 110 to 430 |imol/m2/day. were at times greater than NOs
flux into the sediments in the St. Lawrence estuary and Gulf of St. Lawrence. Many
species of algae and bacteria will take up DON; thus, this flux of DON out of the
sediments leads to an important sink for N within the system (Alkhatib et al.. 2013).
Reactivity of organic matter is another key factor in DON availability.
7.2.6.3.	Dissolved Inorganic Nitrogen
Dissolved inorganic nitrogen (NH4+ plus INO2 + NO3 |) plays an important role in N
cycling and primary productivity. N laden organic matter may be transported to estuaries
and coastal waters from river basins. Estuarine benthic communities degrade this
allocthonous organic matter, releasing NH44" at the aerobic sediment-water interface, and
NH4+ further enhances estuarine productivity via oxidation to NO3 . Generally, NH4+ is
considered the preferred form for some phytoplankton, including harmful species, due to
lower energy requirements for uptake (Glibert et al.. 2016) (Appendix 10.2.2 and
Figure 10-7). Remineralization of organic matter is sensitive to redox conditions (Reddv
and Patrick. 1984). sediment grain size (Nowicki and Nixon. 1985). and temperature
June 2018
7-84
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
(Kemp and Bovnton. 1984). If NH4 has to compete with saltwater cations for adsorption
sites, then excess NH44" becomes free in these saline waters (Gardner etal.. 1991).
The understanding of DIN's role in estuarine and near-coastal nutrient enrichment is an
active field of research. Cornwell et al. (2014) measured September 2011 versus March
2012 benthic nutrient fluxes in the San Francisco Bay and delta and ascertained that
sediment DIN is an important source of N to the water column and system productivity.
Paudel et al. (2017) examined the effect of flow regimes on DIN release at the
sediment-water interface. They compared two south Texas estuaries: the Nueces and the
Guadalupe. The Guadalupe has eight times the inflow of the Nueces. They observed
significantly different NH44" concentrations between the estuaries; however, the
NO2 + N03 concentration did not differ significantly. Using the coastal water
phytoplankton Redfield ratio (ratio of N:P in phytoplankton) of 16:1, they found the
Nueces N:P ratio below the Redfield but the Guadalupe above, suggesting that the
Nueces was N limited. NH44" releases were greater in the Nueces estuary due to its higher
organic content in sediment. In the Guadalupe estuary, finer sediment particles may retain
N.
Buzzelli et al. (2013b) applied the Land-Ocean Interactions in the Coastal Zone (LOICA)
approach to gain understanding about the seasonal budgets for DIN in two south Florida
estuaries—Caloosahatchee River estuary (CRE) and the St. Lucie estuary (SLE)—during
the years 2002 to 2008. The analysis included the contribution of direct wet atmospheric
N deposition to estuarine loading. They compared the production of C, N, and P in the
two estuaries and source attribution. These two watersheds are highly disturbed due to
urban growth and agriculture and flush at substantially different rates. They sought to
understand the optimum metabolism of each estuary from a C, N, and P production and
consumption perspective to better inform coastal watershed management planning. Both
estuaries have similar agricultural and urban land use; however, study authors observed
increased DIN in the CRE during the wet season but not in the SLE. The SLE's
"muck-like" sediment reduces light penetration and isolates the water column from the
benthic zone. (Buzzelli et al.. 2013a: Sime. 2005). External loading influenced the SLE's
production of C, N, and P more than the CRE. Even though the CRE is 2.5 times larger
(and receives double the freshwater inflow) as compared with the SLE, the CRE's DIP
and DIN loadings are only 60-70% of SLE's loadings when spatially normalized.
7.2.6.4. Nitrification
Nitrification is the microbially mediated conversion of NH4 to NO3 : thus, exchanging
one bioavailable N form for another. The process includes NH3 oxidation to NO2 .
June 2018
7-85
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
followed by NO: oxidation to NOs . This importance was well known at the time of
preparation of the 2008 ISA. More recent research has provided additional quantitative
context, including new studies in North America and Europe. Where DIN loading is not
dominated by NO3 . the oxidation of NH44" to form NO3 largely controls the relative
abundance of oxidized and reduced DIN in estuaries. Nitrification rates are often
correlated with suspended particulate matter and NH44" concentration in water (Damashek
et al.. 2016).
There have been new findings about the process of nitrification since the 2008 ISA. It has
generally been assumed that the first step of nitrification (NH4+ oxidation to NO2 ) is the
rate-limiting step (Damashek et al.. 2015). However, recent research indicates that
physical factors such as salinity, temperature, pH, DO concentration, and light can affect
the rates to the extent that the two steps of nitrification may not always be "coupled" as is
often assumed (Hciss and Fulweiler. 2016; Bristow et al.. 2015). In some cases, the rate
for the second step of the nitrification process may be higher than for the first step. For
example, results showed that rates of NO2 oxidation to NO3 were negatively correlated
with light and pH, indicating in part that NO2 oxidation rates are higher when the pH is
lower (Hciss and Fulweiler. 2016). Decoupling of the two nitrification steps was also
observed in the Gulf of Mexico hypoxic zone, where the first step occurred at rates of up
to 30 times higher than the second step, a result the authors concluded was due to
environmental factors such as temperature, substrate availability, and hypoxic conditions
(Bristow et al.. 2015). These results add complexity to the understanding of nitrification
rate and the degree to which that process is affected by environmental factors, and
support the need to measure the two steps of nitrification separately to better predict the
effects of perturbations to the N cycle in coastal waters.
Nitrification makes denitrification possible to some degree, and the reactions are often
assumed to be coupled, as nitrification provides the NO3 starting point for the N
reduction process of denitrification. Since the 2008 ISA, the degree of coupling (and
"decoupling") between nitrification and denitrification reactions has been the subject of
much research (see discussion in Appendix 7.2.6.8).
7.2.6.5. Dissimilatory Nitrate Reduction to Ammonium
DNRA has been characterized as an important pathway in which N is conserved in
coastal ecosystems (Giblin et al.. 2013). Although DNRA was not discussed in the 2008
ISA, more recent research has established DNRA as a potentially important N reduction
pathway, and one that varies in magnitude depending on a range of environmental
factors. Ammonium produced via DNRA can enhance productivity and respiration,
June 2018
7-86
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
which in turn may exacerbate hypoxia (McCarthy et al.. 2015). Rates of DNRA usually
vary seasonally and peak in the warmer summer months. Bernard et al. (2015) found that
in Little Lagoon, AL, DNRA accounted for up to 30-40% of the annual NH4 flux, and
up to 80% of the seasonal NH4 flux during summer months. Also in the summer, DNRA
rates (study average: 52.1 (.iniol N/m2/h) outpaced the removal of NOs via conversion to
N2 (denitrification) (study average 7.7 (.iniol N/m2/h), resulting in more bioavailable N
retained in the system. This could help promote eutrophic conditions (Bernard et al..
2015).
Several reasons have been proposed for observed seasonal peaks in DNRA rates. These
include warmer temperatures, higher sulfide (HS ) and oxygen concentrations, and more
N03 availability. Higher HS concentrations are shown to favor DNRA rather than
denitrification (Kraft et al.. 2014; Howarth et al.. 2011). In the Niantic River estuary, CT,
one study area with the highest rates of DNRA was also found to have high HS
concentrations and reaction rates of sediment denitrification, anaerobic oxidation of
ammonium (Appendix 7.2.6.7). and DNRA varied considerably within the estuary.
Approximately one-third of the total area of the estuary exhibited rates of DNRA that
exceeded denitrification by at least 20%, although denitrification accounted for about
90% of NO, reduction across the entire estuary (Plummer et al.. 2015). Sulfide was
found to predict 44% of the variability in a DNRA ratio metric, while organic carbon
abundance and organic carbon source were less strongly correlated with DNRA
(Plummer et al.. 2015). Declevre et al. (2015) observed wide variations in N removal
pathway rates over small distances in the Paulina polder mudflat (Westerschelde estuary,
Netherlands). Rates of DNRA varied significantly on a small scale of less than 2 m and
were significantly related to HS production.
Jantti and Hietanen (2012) demonstrated that DNRA controlled the overall NO3
reduction under conditions of low oxygen. DNRA is thought to provide energy to
diatoms in dark and/or hypoxic conditions (Glibert etal.. 2016; Kamp etal.. 2011). A
study from Australia's Yarra River estuary found that DNRA rates were increased under
oxygen saturation and depressed under hypoxic conditions (Roberts et al.. 2014). This
result may be explained by increased iron availability and the binding of Fe2+ with free
sulfides. Increased iron availability in sediments has been linked with increased rates of
DNRA (Robertson et al.. 2016).
7.2.6.6. Denitrification
Bioavailable N removal from estuarine and near-coastal ecosystems occurs via several
different pathways in the N cycle. These pathways include sediment burial, uptake by
June 2018
7-87
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
plants, denitrification (NOs reduction to N2) and anaerobic NH44" oxidation to N2 (termed
anammox; Appendix 7.2.6.7V In some areas, uptake by vegetation such as SAV can
represent a significant N sink (Zarnoch et al.. 2017; Bovnton et al.. 2014; Havn ct al..
2014). However, much of the N contributed to riverine estuaries by atmospheric
deposition and other nonpoint and point sources of N is removed from the aquatic
ecosystem by either denitrification or anammox (Damashek et al.. 2015; Ward. 2013;
Bovnton and Kemp. 2008). These processes predominate in the anoxic sediments if the
water body is well mixed and occur in the water column of stratified hypoxic systems.
Since the 2008 ISA, there have been additional insights into factors that affect
abundance, biodiversity, and biological activity of denitrifying microbial communities
(Appendix 7.2.6.9).
Denitrification can occur in both the hypoxic water column and hypoxic sediments,
although in most estuaries sediment processing probably exceeds water column
processing (Seitzinger et al.. 2006). Denitrification is a very important sink for N in most
estuaries (Havn et al.. 2014; Nixon et al.. 1996). Because N is such an important limiting
factor for primary production in estuaries, the removal of N through denitrification is a
valuable ecosystem service in terms of constraining the extent and magnitude of
eutrophication (Piehler and Smyth. 2011 Smyth. 2013. 2481620). In the eutrophic Baltic
Sea, denitrification in sediments is important for partially mitigating the adverse effects
of eutrophication (Jantti and Hietanen. 2012). However, denitrification rates are not
always high enough to result in a net sink of N within an estuary or embayment. Barnes
and Upstill-Goddard (2011) reported measurements of dissolved nitrous oxide (N2O),
inorganic N, O2, and turbidity in six estuaries in the U.K. and results suggested that the
main source of N2O was nitrification; denitrification did not appear to be a significant
N03 sink in that ecosystem.
Denitrification was found to be the primary pathway for N reduction in the Niantic River
estuary in Connecticut, accounting for an average of 90% of total N reduction, a result
that is in line with measurements from other temperate estuaries (Plummer et al.. 2015).
Estuaries and coastal embayments with longer water residence times and shallower water
depths have also been found to exhibit greater rates of denitrification (Havn et al.. 2014).
Shallow waters enhance denitrification because they lead to a greater level of interaction
between N in the water and sediments (Havn et al.. 2014; Nixon etal.. 1996). An analysis
of estuary nutrient budgets for North Atlantic estuaries by Nixon et al. (1996) suggested
that the fractional transport of nutrients through estuaries to the continental shelf is
inversely correlated with water residence time; specifically, the fraction of the total
export of N from the estuary is roughly proportional to the log mean residence time of the
water in the estuary. Havn et al. (2014) studied the nutrient dynamics of West Falmouth
Harbor on Cape Cod, MA, a shallow estuary that experienced a large increase in N load
7-88
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
from human activities without a substantial change in P load. Since the 1990s, this
estuary received a threefold increase in N inputs caused by groundwater contamination
by a municipal wastewater treatment plant. During summer, the Falmouth Harbor
retained most of the N load contribution and also imported some additional N from the
adjacent Buzzards Bay; during spring and fall, N was exported from the Harbor to the
Bay.
Inhibition of denitrification rates by HS has been widely observed. Plummer et al.
(2015) found that denitrification rates were inversely related to pore water HS (as a
measure of SO42 reduction), consistent with previous studies. Organic carbon abundance
and organic carbon source were also shown to be related to denitrification rates in that
study (Plummer et al.. 2015).
Smyth et al. (2013) measured the N2 fluxes in five major estuarine habitat types in Bogue
Sound, southeastern North Carolina: salt marshes, seagrass beds, oyster reefs, and inter
tidal and subtidal flats. Based on the distribution of habitats across the study estuary, it
was estimated that about 76% of the watershed N load was removed by denitrification in
the estuary each year. Results suggested that the amount of N removed by denitrification
from an estuary depends on the amount and type of habitats located in that estuary. Each
habitat had a characteristic impact on N cycling in the sediment (Smith et al..
2013).Smyth et al. (2013); Piehler and Smyth (2011) concluded that more structured
habitats such as oyster reefs and seagrass beds provided especially high N removal per
unit estuary area. Calculations showed that the SAV habitat produced the highest value of
N removal in terms of an ecosystem service valued at $3,000/ac/year, compared to
$400/ac/year" value provided by a subtidal flat habitat (Piehler and Smyth. 2011).
7.2.6.7. Anammox
Anammox removes N from estuaries by producing N2 gas. This process can in some
cases be a significant pathway for N removal, depending on environmental factors,
although it is typically secondary to denitrification in N removal. Anammox and
denitrification are usually highly correlated (Plummer et al.. 2015; Lisa etal.. 2014). A
review of several studies revealed that anammox rates in coastal sediments can
potentially range up to 52 jjxnol N/m2/h (Plummer et al.. 2015). McCarthy et al. (2015)
found that up to 29% of the N removal (mean = 11.8 ± 1.7%) from the Mississippi River
TN load to the Louisiana-Texas continental shelf may be due to anammox under hypoxic
conditions. The median denitrification rate measured in that study, which includes
anammox, of 88.1 |imol N/m2/h was found to be comparable to other studies (McCarthy
et al.. 2015). In the Niantic River estuary, CT, anammox rates ranged from 0 to 3.1 |imol
June 2018
7-89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
N/m2/h, accounting for 3.5% of total N reduction, while denitrification accounted for
91% of total N reduction (Plummcr et al.. 2015). Anammox was also found to be a
significant N sink in the lower St. Lawrence estuary, where it was measured at a rate of
5.5 ± 1.7 |imol N/m2/h (Crowe et al.. 2012). In the Cape Fear estuary, NC, anammox
rates ranged from 0.17 to 4.77 nmol N/g wet sediment/h, with the highest rates measured
during the winter (Lisa et al.. 2015). Anammox rates in this study were also found to be
inversely related to temperature.
Several studies have linked anammox activity to N loading. In the Chesapeake Bay,
higher anammox activity was associated with higher concentrations of NOs in the tidal
freshwater segment and was not observed in the lower saline part of the estuary (Rich et
al.. 2008).
7.2.6.8. Nitrification/Denitrification Uncoupling
Sediment N cycling is dynamic, and rates of various processes fluctuate over time.
Recent research has highlighted the importance of quantifying each process, the resulting
N fluxes, and environmental controls on these reactions. Denitrification has long been
assumed to be coupled with nitrification; however, the extent of that coupling is
increasingly understood to be dependent on environmental factors. Hines etal. (2012)
found that 43% of denitrification was coupled to nitrification in the upper Cape Fear
River estuary, NC. Lisa et al. (2015) found that nitrification and denitrification rates were
significantly coupled (correlated) in the Cape Fear River estuary, and that this coupling
was driven by organic carbon mineralization. The study found that tidal changes and
salinity fluctuations did not affect the coupling of nitrification/denitrification on the same
scale (Lisa et al.. 2015). Under some conditions, such as warmer summer conditions,
DNRA may be the favored N reduction pathway over denitrification, which can also lead
to uncoupling of nitrification/denitrification rates. In Little Lagoon, AL, DNRA rates can
be much higher in the summer and are correlated with higher concentrations of HS
(Bernard et al.. 2015). Seasonality was also a factor in the San Francisco Bay, where
coupling of nitrification/denitrification was more prominent in the late summer; in the
spring, high NOs concentrations in the water drove denitrification (Cornwell et al..
2014).
June 2018
7-90
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
7.2.6.9. New Insights Regarding the Abundance, Biodiversity, and Biological
Activity of Estuarine Microbial Communities
Since the 2008 ISA, there is new information on the abundance, diversity
(Appendix 10.3.4) and biological activity of microbial communities and their role in N
cycling. Quantification of functional genes reflect changes in microbial function and have
been linked to forms of N. For example, in seasonally anoxic bottom waters in
Chesapeake Bay, Eggleston et al. (2015) observed that relative expression of genes
involved in denitrification and DNRA were coincident with changes in concentrations of
NOs . NO2 . andNH4+ In an analysis of denitrifiers in San Francisco Bay estuary, N was
identified as one of the key factors (along with salinity, organic C and several metals)
affecting community structure and function and denitrification rates (Mosier and Francis.
2010). Seasonal inputs of N loading from the watershed affect microbial community
structure and biodiversity, which in turn affect rates of processes involved in N cycling
(Lisa et al.. 2015; Lisa etal.. 2014).
7.2.6.10. Archaea and Nitrogen Cycling
Prior to the 2000s, it was generally believed that NH3 oxidation was accomplished only
by Proteobacteria (Damashek et al.. 2015; Kowalchuk and Stephen. 2001). Complexity
increased with the discovery that some Archaea (also called Thaumarchaeota), a
primitive life form separate from bacteria, can also oxidize NH3 (Brochier-Armanet et al..
2008; Konneke et al.. 2005). These ammonia-oxidizing Archaea (AOA) are dominant in
some estuaries (Moin et al.. 2009). Ammonia-oxidizing bacteria (AOB) are important in
others (Abell et al.. 2010). These two life forms can vary spatially and seasonally in
importance in still other estuaries (Damashek et al.. 2015; Urakawa et al.. 2014; Zheng et
al.. 2014). Benthic NH3 oxidizers may be able to oxidize a significant quantity of NH4 in
NIL+ rich systems, such as the Sacramento River (Damashek et al.. 2015). Several
studies that have examined changes in AOA and AOB relative abundance and
community structure associated with N loading are reviewed in Appendix 10.3.4.
7.2.6.11. Role of Benthic Macrofauna in Nitrogen Cycling
The role of benthic macrofauna (including crustaceans and molluscs) in N and C cycling
and their ability to modulate water quality have significant implications for estuarine
functioning (Rose et al.. 2015a; Bricker et al.. 2014; Petersen et al.. 2014; Rose et al..
2014; STAC. 2013; Carmichael et al.. 2012; Volkenborn et al.. 2012; DAndrea and
DeWitt. 2009; Cerco and Noel. 2007). Activities of burrowing macrofauna can create
June 2018
7-91
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
areas of oxic-anoxic oscillations, which vary on the order of minutes to hours and affect
geochemical reactions and microbial activity in sediments (Volkenborn et al.. 2012).
Burrowing mudshrimp (Upogebia pugettensis) were shown to increase the rate of N
cycling processes and DIN fluxes in an intertidal mud flat in the Yaquina River estuary in
Oregon (D'Andrea and DeWitt. 2009). Stief (2013) reviewed the contributions of benthic
macrofauna to the turnover of N and to emissions of greenhouse gasses such as nitrous
oxide from the estuary to the atmosphere. Sediment burrowing macrofauna stimulated
nitrification and denitrification in the sediment. Together these facilitate removal ofN
from the estuary system. Benthic macrofauna intensify the coupling among water,
benthos, and atmosphere by enhancing turnover and transport of N.
The use of shellfish for remediation of coastal N enrichment has been explored due to the
ability of these organisms to modulate nutrient dynamics and water quality (Rcitsma et
al.. 2017; Ferreira and Bricker. 2016; Petersen et al.. 2016; Bricker et al.. 2014; STAC.
2013; Cerco and Noel. 2007; Carmichael et al.. 2004). These filter feeders store nutrients
in shell and tissue that are permanently removed from the estuary with shellfish harvest.
Biodeposits (partially digested phytoplankton expelled from suspension feeding bivalves)
may remove additional N through acceleration of denitrification processes in underlying
sediments (Pollack et al.. 2013; STAC. 2013; Stephenson et al.. 2010). For example, in
the Mission-Aransas estuary in Texas, oyster reefs were estimated to remove 502.5 kg
N/km2 through denitrification of biodeposits and 251.3 kg N/km2 in burial of biodeposits
to sediments (Pollack et al.. 2013). Oyster harvest in the same estuary was calculated to
remove approximately 21,665 kg N/year.
Additional studies have reported N removal by shellfish in U.S. waters. Carmichael et al.
(2004) directly measured N removal in the oyster C. virginica in transplant studies to five
Cape Cod estuaries with different N loadings. Estimated N removal of different sources
was <15% of land-derived N loads and <1% of phytoplankton. Nitrogen removal by
oysters in the Great Wicomico River, which drains into the Chesapeake Bay, was
estimated to be 15.2 tons/year [total area 2.8 x 105m2; (Cerco. 2015)1. Based on model
estimates for Chesapeake Bay harvesting 7.7 * 106 harvest-sized oysters (76-mm)
removes 1 ton of N from the bay (Higgins etal.. 2011). In the same study, an offset of
approximately 10 to 15% of total N load in some basins (cultivation of
200 x 106 oysters/year) was calculated. Holmer et al. (2015) recommended harvest of
mussels within 1 year of the production cycle for the most efficient N removal.
Over time, shellfish excretion and subsequent sedimentation may contribute to
development of hypoxic conditions. A shellfish farming operation may shift from a net
sink to a net source of N. Oyster reef restoration projects in several locations throughout
the U.S. have been evaluated for effects on N cycling; however, this process appears to
7-92
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
be dependent on local habitat and growing conditions and may vary by orders of
magnitude (Cerco. 2015; Smyth et al.. 2015; Kellogg et al.. 2014; Plutchak et al.. 2010).
In Great Bay estuary, NH, eutrophication enhanced oyster feeding rates and enhanced
biodeposit quality. Thus, oyster-mediated ecosystem services may be expected to vary
with environmental conditions (Hocllcin et al.. 2015).
Recent evaluation of oysters in Chesapeake Bay for inclusion in TMDL nutrient
reductions reported enhanced denitrification in association with oyster reefs; however,
the effect was highly variable and not reliable unless direct measurements were
conducted on individual reefs (STAC. 2013). The potential for shellfish aquaculture to be
included in proposed nutrient trading markets for achieving pollution control under the
Clean Water Act is being evaluated in Chesapeake Bay (Stephenson et al.. 2010). Some
recent modeling studies showed that mean N removal by shellfish aquaculture compares
favorably with reported N removal effectiveness of agricultural best management
practices and stormwater control measures (Rose et al.. 2015a; Rose et al.. 2014).
7.2.6.12. Climate Modification of Ecosystem Response to Nitrogen
Altered biogeochemical processes due to N loading are occurring within the context of
climate change. Microbial diversity is affected by environmental gradients within
estuaries, a condition that may be exacerbated under climate change. For example,
freshwater and nutrient contributions to estuaries are expected to rise due to predicted
increases in surface water flow and runoff from watersheds (Rabalais et al.. 2010; Adrian
et al.. 2009; Whitehead et al.. 2009). Howarth et al. (2012) demonstrated larger N fluxes
(larger percentage delivery of human N inputs) in wetter climates with more discharge,
across 154 different watersheds in the U.S. and Europe. Temperature modification
leading to sea level rise and inputs of fresh water will likely influence the delivery of
nutrients and organic matter, alter salinity gradients, and increase stratification within
estuaries (Statham. 2012). Eutrophic conditions and the extent and duration of hypoxia
are predicted to increase with anticipated changes in temperature and precipitation
(Altieri and Gedan. 2015; Rabalais et al.. 2009; Boesch et al.. 2007). High organic loads
and freshwater inputs associated with extreme weather events may further enhance
thermal stratification and contribute to hypoxia (Wetz and Yoskowitz. 2013). Increased
thermal stratification will worsen hypoxia where it already occurs and may facilitate its
formation at other locations (Rabalais et al.. 2010).
The "benthic filter" is a term given to the benthic microbe and algal communities that
remove N via uptake and denitrification and bury N from shallow estuarine and coastal
waters. Nitrogen removal via the benthic filter is an important mediator of nutrient
June 2018
7-93
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
enrichment in shallow estuaries along the Atlantic and Gulf coasts of the U.S. where this
feature is common (Anderson et al.. 2014b). The filtering function of this top layer of
sediment is affected by nutrient delivery as well as light availability, temperature, and
resuspension caused by wind and storms. Thus, the benthic filter and the N removal
services it provides are expected to be highly susceptible to climate change impacts
(Anderson et al.. 2014b').
7.2.7. Monitoring Data
Monitoring data provide temporal trends of biogeochemical processes and indicators
associated with eutrophication and nutrient-enhanced coastal acidification in estuaries
and coastal ecosystems. These data document and quantify changes that occur in
response to environmental stressors, including N deposition. Many monitoring studies for
eutrophication have been ongoing for one or two decades, in some cases longer.
Monitoring for coastal acidification is more recent. Analysis of historical data suggest
that some coastal areas have been experiencing long-term acidification.
7.2.7.1. Eutrophication
Since the 2008 ISA, nearly a decade's worth of additional data have been added to some
of the monitoring programs described below. The availability of these additional data
facilitates trend detection now, compared to 2008. The U.S. EPA Estuary Data Mapper
tool accessed at https://www.epa.gov/hesc/about-estuarv-data-mapper-edm allows users
to retrieve and visualize estuary data from several federal agencies to access water and
sediment quality, freshwater discharge, tides, N deposition, and other parameters.
The NARS include a survey of coastal waters called the National Coastal Condition
Assessment (NCCA). U.S. EPA, together with the states, tribes, other entities, and
individuals, have collaborated on these statistically representative surveys since the early
2000s. The NCCA survey uses standardized field protocols and indicators of coastal
condition including N, water clarity, chlorophyll a, and DO concentrations. The NCCA
2010 used a consistent set of data from three periods (1999-2001, 2005-2006, and 2010)
to evaluate change in coastal conditions overtime (U.S. EPA. 2016g). This analysis
included only the Northeast, Southeast, Gulf, and West Coast regions. (The change
analysis does not include the Great Lakes because they were not part of this survey until
2010.) The change analysis showed that scores on the water quality index remained
relatively similar between 2005-2006 and 2010, after a significant decrease in the
percentage of area rated good from 1999-2000 to 2005-2006. The Gulf Coast had the
June 2018
7-94
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
greatest proportion of waters (24.4%) rated poor for water quality compared to the other
regions surveyed for the NCCA in 2010 (Figure 7-9). The NCCA report noted that these
findings support the need for continued attention to coastal stressors at national, regional,
state, and watershed scales to identify and mitigate challenges where they exist and
protect areas that are still in good condition (U.S. EPA. 2016g).
Source: U.S. EPA (2016a).
Figure 7-9 Percentage of area in each coastal region scoring good, fair, and
poor based on the Water Quality Index for the NCCA 2010.
Additional long-term monitoring surveys conducted with help from the U.S. EPA include
those undertaken by individual National Estuary Programs (NEP) around the U.S. The
NEP is a nonregulatory program designed to protect and restore the water quality and
ecological integrity of estuaries of national significance, of which there are currently 28.
The NEPs develop and implement long-term management plans that contain actions to
address water quality priorities, which are defined by local, city, state, federal, private,
and nonprofit stakeholders. In many cases these priorities include reducing and mitigating
the occurrence of eutrophic conditions in large estuaries, goals often accompanied by
long-term monitoring of nutrients and eutrophication indicators. For instance, the Tampa
Bay 'Estuary Program works with the Southwest Florida Water Management District's
June 2018
7-95
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Surface Water Improvement and Management Program to collect monitoring data on
trends in seagrass coverage, which is greatly affected by water clarity (often an indicator
of excessive N loads to the bay) (Sherwood. 2017). The most recent analysis of
monitoring data showed that Tampa Bay's seagrass area continues to recover, with the
addition of 1,360 acres of seagrass coverage reported from 2014 to 2016. Tampa Bay's
total seagrass coverage is now estimated to be 41,655 acres as of 2016, which now
exceeds the total estimated seagrass coverage in the 1950s period (40,420 acres) for the
first time since monitoring began (Sherwood. 2017). See the Tampa Bay Case Study
(Appendix 16) for additional information on this coastal system.
The NEP's Long Island Sound Study has conducted two decades of monitoring to
identify N sources and track water quality conditions and eutrophic indicators over time.
The Long Island Sound Study uses several indicators, including the NCCA's Water
Quality Index, to evaluate changes from year to year. Data from 1991-2011 show a
gradient in improving water quality from the eastern part of the sound, where population
and anthropogenic stressors are lower and water quality was most often rated "good," to
the western side where population and development pressure are both higher and water
quality was consistently rated "fair" (LISS. 2017).
The Chesapeake Bay Monitoring Program established in 1984 is a regional partnership
that monitors 19 physical, chemical, and biological characteristics, including nutrients
and DO, enabling the study of the bay and long-term trends (Testa et al.. 2017). The
original Chesapeake Bay Program agreement of 1983 included Maryland, Virginia,
Pennsylvania, Washington, D.C., the U.S. EPA, and regional partners. In 2014, the
Chesapeake Watershed Agreement was signed to include representatives from the entire
watershed and accelerate restoration of the bay.
The NPS also conducts estuarine water quality monitoring to monitor "Park Vital Signs,"
which are elements and processes of park ecosystems that can serve as indicators of the
overall health of the park. The Estuarine Nutrient Enrichment Monitoring (ENEM)
program monitors nitrogen loading inputs as well as overall water quality and seagrass
distribution in park estuaries within the Northeast Coastal and Barrier Network (NCBN)
region (covering Massachusetts to Virginia) of the NPS. In this region, the Park Vital
Signs are all related to nutrient enrichment, which is considered the stressor with the
ability to cause the greatest potential impacts on park health (USGS US Department of
the Interior. 2015). Water quality indicators are defined with the same criteria as U.S.
EPA's NCCA methods, and data are available on the program's website for each park in
the NCBN region of the NPS at
https ://science .nature .nps. gov/im/units/ncbn/monitor/estuaries .cfm.
7-96
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
The National Oceanic and Atmospheric Administration (NOAA) conducts several
monitoring programs in estuaries and coastal waters. The National Estuarine Research
Reserves System (NERRS) is a federal-state partnership network of 29 coastal sites
encompassing 1.3 million acres designated to protect and study estuarine systems.
Estuarine water quality monitoring data are collected by the NERRS System-Wide
Monitoring Program (SWMP), which identifies and tracks short-term variability and
long-term changes in estuarine ecosystems and coastal watersheds. Data are housed in a
central online database available at http://cdmo.baruch.sc.edu/.
Many NERRS sites work with other groups to analyze and report their overall findings
about water quality indicators. In the Great Bay estuary in New Hampshire and Maine,
NERR's SWMP collects nitrogen, DO, and temperature data, which is used by the
Piscataqua Region Estuaries Partnership (an NEP site) to assess the environmental status
and trends in the Bay. The most recent State of the Estuaries report found that total N
load to the estuary in 2009-2011 was 1,225 tons per year, which appeared to be strongly
influenced by rainfall amounts (PREP. 2013).
NOAA's Gulf of Mexico Ecosystems and Hypoxia Assessment (NGOMEX) is a large
study focusing on a single geographic region in the Gulf of Mexico. NGOMEX has been
collecting data and conducting research for more than 30 years to investigate the spatial
and temporal dynamics of the hypoxia zone in the Gulf of Mexico, especially with
regards to the link between the hypoxia zone size and the amount of nutrient loading,
primarily from the Mississippi River watershed (Appendix 10.2.4V In 2017, work funded
by NGOMEX measured the largest Gulf of Mexico hypoxia zone since mapping of the
area began in 1985 (U.S. EPA. 20176). The size of the 2017 hypoxia zone was accurately
forecasted by Mississippi River spring discharge levels and nutrient data, gathered by the
USGS and analyzed with NOAA-sponsored models, and clearly indicates that nutrient
pollution from the Mississippi River watershed is affecting the health of the coastal Gulf
ecosystem.
NOAA also conducts the Monitoring and Event Response for Harmful Algal Blooms
(MERHAB) program which helps identify when beaches, shellfisheries, and marine
animals are at risk from harmful algae and cyanobacteria. It allows local stakeholders to
react as quickly as possible to any human health risk. MERHAB largely funds research
projects, but also aims to help routine water quality and shellfish monitoring studies
upgrade to better HAB detection methods and technologies. NOAA's Center for
Operational Oceanographic Products and Services (CO-OPS) produces HAB forecasts
for the Gulf of Mexico to help communities plan for and mitigate the effects of
potentially harmful algal blooms. Data for these forecasts come from many sources. The
National Aeronautics and Space Administration (NASA) Moderate Resolution Imaging
7-97
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Spectroradiometer (MODIS) satellite collects ocean color data which are processed by
NOAA CoastWatch (NOAA. 2017a). The satellite sensors measure visible light at
specific wavelengths to determine the color of the ocean. These color data can be used to
estimate chlorophyll concentrations; however, in coastal areas, interpretation can be
complicated by the presence of other biota, compounds, and minerals (NOAA. 2017b).
The HAB forecasts prepared by CO-OPS also integrate other data about factors that may
affect algal and cyanobacterial bloom formation and intensity such as water temperature,
ocean currents, and weather conditions (NOAA. 2017b). The forecasts also note cell
counts of the toxic dinoflagellate Karenia brevis (commonly called a "red tide" species)
taken from specific locations along the coast. Another NOAA product called the Harmful
Algal Bloom Observing System (HABSOS) integrates these data sources, including cell
counts and other environmental information into an interactive mapping application
available online at https://habsos.noaa.gov/ (NOAA. 2017c). HABSOS can map recent
satellite data and cell counts to depict areas that may be at risk of HABs in the near
future. Trends in HABs and responses to N are discussed further in Appendix 10.2.2 and
Appendix 10.3.3.
7.2.7.2. Coastal Acidification
Coastal acidification programs are developing around the world. In the U.S., ocean
acidification has been documented from the New England region, California, Oregon,
Washington, and the northern Gulf of Mexico (Laurent et al.. 2017; Gledhill et al.. 2015;
Gruberetal.. 2012; Hauri et al.. 2009; Feelv et al.. 2008; Yang. 1998). In a series of
sampling cruises and analysis of water sampling data, Wallace et al. (2014) identified
concurrent low pH conditions and declining DO in four Northeast estuaries (Long Island
Sound, NY; Jamaica Bay, NY; Narragansett Bay, RI; Hempstead Bay, NY). Observed
conditions in eutrophic estuaries during late summer were such that marine biota,
especially calcifying organisms, may be affected (Appendix 10.5). Analysis of historical
(from the late 1960s to 2010) alkalinity and pH data from bays along the northwestern
coast of Texas indicated that most of the bays in this region have experienced long-term
acidification (Hu et al.. 2015a).
7.2.8. Modeling Estuaries and Near-Coastal Areas
Since the 2008 ISA, several new applications of existing models have quantified
eutrophication processes in estuaries and near-coastal marine ecosystems. These have
included studies that focused primarily on N cycling or hypoxia.
June 2018
7-98
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
7.2.8.1. Models
There are several models that track the sources and movement of N through the landscape
and stream network to estuary and near-coastal marine ecosystems. These models vary in
scope, level of spatial and temporal resolution, forms of N considered, complexity, and
empirical versus deterministic construction (Alexander et al.. 2008). They can be
structured as statistical models (Howarth et al.. 2012; Peierls et al.. 1991). empirical
models (C'araco and Cole. 1999). export coefficient models (Johncs. 1996). deterministic
models such as SWAT (Gassman et al.. 2007; Nietsch et al.. 2002). Generalized
Watershed Loading Function [GWLF; (Haith and Shoenaker. 1987)1. and hybrid
approaches (SPARROW, NEWS). The principal ecosystem models applied to U.S.
waters to assess N enrichment were reviewed in the 2008 ISA and are briefly summarized
here (U.S. EPA. 2008a).
Some of these models have been applied at national or large regional scales to examine N
loading to rivers and the coastal zone. SPARROW is a hybrid statistical-mechanistic
model that can attribute pollutant sources and contaminant transport to surface waters
(Smith et al.. 1997). SWAT is a mechanistic model developed by the U.S. Department of
Agriculture, Agricultural Research Service (Gassman et al.. 2007; Nietsch et al.. 2002).
WATERS-N is a mass balance model that has been used to evaluate N inputs to estuaries
(Castro et al.. 2003; Castro et al.. 2001).
Since the 2008 ISA, there have been new applications of SPARROW and SWAT, as well
as development and application of new models and approaches such as NEWS and Net
Anthropogenic Nitrogen Inputs (NANI). SPARROW has been used to estimate total N
loads within watersheds to estimate sources of N to streams and rivers. It has also been
applied at regional and national scales (Schwarz et al.. 2011; Alexander et al.. 2008).
SPARROW operates on an annual time step, usually integrating several decades of data
to develop load relationships and then simulating conditions for 1 year. NEWS is a
similar hybrid model. It has been used to estimate the magnitudes and sources of different
forms of N (particulate, dissolved inorganic, and organic) to coastal waters (Glibert et al..
2010a). Recent work with NEWS included a global, seasonal version (McCrackin et al..
2014). It has been used for scenario comparisons, including predictions about reductions
in air deposition resulting from Clean Air Act regulations and other policy actions
(McCrackin et al.. 2015). NANI is a simple mass balance model for calculating net
anthropogenic nitrogen inputs that takes information about inputs and outputs of nitrogen
within a basin and estimates the riverine flux from the landscape to the ocean (Howarth et
al.. 2012). It was developed for watersheds across the contiguous U.S. at the county level
(Hong et al.. 2013. 2011). Using NANI, Howarth et al. (2012) demonstrated larger N
fluxes (larger percentage delivery of human N inputs) in wetter climates with more
June 2018
7-99
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
discharge, across 154 different watersheds in the U.S. and Europe. The NANI Calculator
Toolbox takes into consideration fertilizer N application, agricultural N fixation, net food
and feed imports, and atmospheric sources of N.
Deterministic models of N flux are based on mechanistic relationships that simulate N
transformations, transport, and removal, often at relatively fine temporal and spatial
resolution. SWAT has been recently applied at the regional scale of the Mississippi River
basin (Santhi et al.. 2014). It allows examination of the effects of changes in cropland
management on delivery of N to coastal waters. Recent SWAT publications do not
explicitly include atmospheric deposition as a source of N, but have produced similar
overall results as SPARROW and NEWS in terms of load and attribution to agriculture in
the Mississippi River basin (-54—61%; White et al.. 2014; McCrackin et al.. 2013).
The watershed models described above route N from the landscape to estuaries and the
coastal zone, but generally do not model the contribution of N directly to the estuary
surface from the atmosphere. These models also deal with the freshwater components of
watersheds and do not predict N fate within estuaries or N effects on estuarine processes
or functions. The following sections describe new applications of several models that
operate within estuaries.
Additional models and tools that have been applied to assess atmospheric contributions of
N to U.S. estuaries since the 2008 ISA included the watershed N Loading Model [NLM;
(Latimer and Charpentier. 2010)1 and the Watershed Deposition Tool [WDT; (Poor et al..
2013b)]. The latter was developed by the U.S. EPA to map atmospheric deposition
estimates to watersheds using wet and dry deposition data from the Community
Multiscale Air Quality Model [CMAQ; (Schwede et al.. 2009)1. This tool links air and
water quality modeling data for use in TMDL determinations and analysis of
nonpoint-source impacts. The NLM has been used for New England estuaries to estimate
total N loading and the relative contributions of the various N sources (Table 7-9) and
develop load response relationships among N inputs and seagrass extent (Latimer and
Rego. 2010).
7.2.8.2. Predicted Response to Nitrogen Loading
Many of the models that estimate N loads to the coastal zone from the landscape and
freshwater inflow have been compared, and there is a good deal of knowledge about their
limitations and uncertainties (McCrackin et al . 2013: Alexander et al.. 2008). In a 2000
National Research Council review, it was determined that these models are
hydrodynamically complex and tend to be specific to particular sites. Thus, they are
difficult to apply broadly (NRC. 2000).
June 2018
7-100
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
The range of coastal ecosystem responses to changing N loads reflect changes in DO,
productivity, SAV cover, and impacts on other organisms. Water residence time can
influence the response of estuaries to nutrient loading because of the effects of flushing
on nutrients, water temperature, plankton, and light penetration. Capturing these
dynamics in a model is challenging (Swancv et al.. 2008).
A recent review of hypoxia modeling described trade-offs between details of complex
models and the need for simpler, more broadly applicable models (Peiia et al.. 2010).
Greene et al. (2009) developed a set of multiple regression models to simulate
relationships between river loads and concentrations of N and P compared with the extent
of hypoxic bottom water. A Bayesian approach was used in combination with a model of
hypoxia to project the O2 demand and extent of hypoxia (Liu et al.. 2010). Estuarine
models have been important to policy decision making and formulation of nutrient
reduction goals (Scavia et al.. 2006; Scavia et al.. 2004). There is an increased interest in
developing coupled models that connect physical-chemical-biological nutrient loading
and fate models (Fennel etal.. 2011) and that connect chemical eutrophication models to
fisheries response models (Cerco et al.. 2010).
The responses of estuaries to N loading is partly determined by environmental
characteristics. Nutrient-phytoplankton-zooplankton (NPZ) models examine biological
responses including the effects of nutrient limitation and zooplankton predation on
phytoplankton dynamics and fish production. The NPZ model described by Swanev et al.
(2008) can assess complex responses of estuaries to different climate variables, land use,
atmospheric loading, and other stressors (Appendix 10.3.8).
New studies have focused on increasing understanding of spatial variation in DIN export
from the watershed to coastal zones. Seasonal patterns of DIN export influence impacts
of coastal eutrophication, including HABs and the development of seasonal hypoxic
zones. McCrackin et al. (2014) predicted seasonal DIN export to coastal regions, using
the NEWS2 model calibrated to measured DIN yield in 77 rivers distributed globally.
The DIN transport efficiency was positively correlated with runoff and negatively
correlated with temperature. McCrackin et al. (2014) concluded that because of landscape
N attenuation, a better representation of land-to-river N transfers, in particular, might
improve models of nutrient cycling.
Eldridgc and Morse (2008) developed a combined water-column and sediment model to
investigate sediment and water-column metabolism and their impacts on development of
hypoxia on the Louisiana shelf. They found that sediment O2 demand was the primary
sink for O2 at the beginning and end of a hypoxic event. Once hypoxia has developed,
however, sediments became isolated from O2 rich upper waters.
7-101
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Key drivers of hypoxia in coastal waters include nutrient loading, mainly from
anthropogenic sources, C supply, and processes that contribute to thermal stratification.
Complex hypoxia models are difficult to validate, however, and this limits the confidence
that can be placed on simulation results. Eldridge and Roelke (2011) recommended the
use of multiple models and identification of similarities in qualitative model behavior.
They provided a brief overview of recent hypoxia models, and identified the need for a
meta-analysis of simulations produced by these models in search of qualitative
similarities. They concluded that more complex models of hypoxia need to be developed
that include three-dimensional hydrology. Sturdivant et al. (2013) developed a model of
hypoxia impact on macrobenthic production in the lower Rappahannock River, a
tributary of the Chesapeake Bay that experiences seasonal hypoxia. Simulation results
suggested that macrobenthic biomass was strongly linked with dissolved O2
concentration. Biomass fluctuations reflected the duration and severity of hypoxia.
Results suggested that hypoxia had a negative effect on biomass, with longer duration
and greater severity resulting in increased loss of biomass.
Glibert et al. (2010a) reviewed modeling approaches to improve understanding of HABs
and their relationship with nutrient inputs. They recommended that predictive capabilities
will likely improve if a suite of modeling approaches is used. This might include
site-specific loading models of nutrient sources and models that couple nutrient discharge
to biological responses.
Dale etal. (2011) investigated biogeochemical processes that affect N cycling in the
sediment of Eckernforde Bay in the southwestern portion of the Baltic Sea, where severe
bottom water hypoxia (and sometimes anoxia) commonly occurs during late summer.
Sediments acted as net sinks for NO3 . of which three-fourths was characterized as
derived from reduction of NO3 to NH4 . The other 25% was characterized as derived
from denitrification of NO3 to NO2 . The NH4 flux was high (1.74 mmol/m2/day), in
response to degradation of organic N at the sediment interface, and was directed out of
the sediment.
7.2.9. National Scale Sensitivity
Characteristics of coastal systems sensitive to eutrophication are described in
Appendix 10.1.4. In the 2008 ISA and the National Estuarine Eutrophication Assessment
(NEEA), a comprehensive survey of eutrophic conditions in the Nation's estuaries
conducted by the National Oceanic and Atmospheric Administration (NOAA), the most
eutrophic estuaries in the U.S. occurred in the mid-Atlantic region, and the estuaries with
the lowest degree of eutrophication were in the North Atlantic rFigure 10-2;(Bricker et
June 2018
7-102
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
al.. 2007)1. Areas of eutrophication-related hypoxia are found on the U.S. East and West
coasts and the Gulf of Mexico (Figure 10-5). During the summer of 2017 the extent of
hypoxia in the Gulf of Mexico was the largest ever recorded in the U.S. (U.S. EPA.
20176). The most recent NCCA report, based on 2010 data and released by U.S. EPA in
2016, analyzed survey data from 1,104 sites, representing 35,400 square miles of U.S.
coastal waters (U.S. EPA. 2016g). In this most recent NCCA report, water quality was
rated good in 36% of coastal and Great Lakes nearshore waters, fair in 48%, and poor in
14%, based on measures of the eutrophication parameters that make up the water quality
index (P, N, water clarity, chlorophyll a, and DO concentrations) (Figure 7-9).
In the U.S., Chesapeake Bay is perhaps the best-documented case study of the effects of
human activities on estuarine eutrophication. Other impacted estuaries identified in the
2008 ISA included Long Island Sound, the Pamlico Estuary in North Carolina, and along
the continental shelf adjacent to the Mississippi and the Atchafalaya River discharges to
the Gulf of Mexico (U.S. EPA. 2008a). In many parts of the U.S., especially the
Southeast, Midwest and Mid-Atlantic deposition of reduced N has increased relative to
oxidized N in last few decades (Appendix 2 and Appendix 10.1.2). Since the 2008 ISA,
additional evidence has shown that reduced forms of atmospheric N play an increasingly
important role in estuarine and coastal eutrophication (Appendix 10.3.3). The form of N
delivered to some coastal areas of the U.S. is shifting from primarily NOs to an increase
in reduced forms of N.
Nutrient-enhanced coastal acidification was not discussed in the 2008 ISA yet the
contribution of increased respiration caused by N enrichment to acidification of coastal
areas has resulted in recent research and monitoring efforts in the U.S.
(Appendix 7.2.7.2). Ocean acidification has been documented from the Gulf of Maine,
California, Oregon, Washington, and the northern Gulf of Mexico (Laurent et al.. 2017;
Gledhill etal.. 2015; Gruber et al.. 2012; Hauri et al.. 2009; Feelv et al.. 2008; Salisbury
et al.. 2008).
Unlike freshwaters where chemical recovery linked to decreased atmospheric deposition
of N and S is observed in some systems (Appendix 7.1.5.1). N inputs from atmospheric
and other sources contribute to the continued water quality degradation in U.S. coastal
waters. There are only a few cases of documented recovery of biological indicators in
U.S. estuaries such as SAV in Tampa Bay (Appendix 16) and Chesapeake Bay
(Figure 10-6).
7-103
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
7.2.10. Water Quality Criteria for Estuaries
As discussed for fresh water in Appendix 7.1.6. the U.S. EPA develops WQC using the
latest scientific knowledge to determine when water is unsafe for people and wildlife.
State and tribal governments may adopt these criteria or use them as guidance in
developing their own. Numeric nutrient criteria are critical tools for protecting and
restoring a water body's designated uses related to N and P nutrient pollution. These
criteria enable effective monitoring of a water body for attaining its designated uses,
facilitate formulation of the National Pollutant Discharge Elimination System (NPDES)
discharge permits, and simplify development of TMDLs for restoring impaired waters.
Generally, it has been considered more complicated to determine appropriate nutrient
criteria for estuaries than for many freshwater systems because estuaries are influenced
by so many variable factors such as tidal fluctuations, salinity gradients, and other widely
varying physical and chemical conditions which impact water quality. The designation of
water bodies within which similar ambient conditions are expected to occur is much more
difficult in estuaries, as the ambient values often fluctuate widely under natural
conditions within the same estuary. Nevertheless, several states have made progress in
the development of numeric nutrient criteria for estuaries. One of the first states to
implement its own criteria for estuaries, Florida designated 62 different estuary segments
and adopted a TN target value for each segment based on extensive monitoring and
modeling studies.
The U.S. EPA is continuing to work with the states to develop numeric nutrient criteria to
better define levels of N that affect U.S. marine and estuarine waters. WQC for TN are
now available statewide in American Samoa, Florida, Hawaii, and the Northern Marianas
Islands (Figure 7-10). Guam has statewide criteria for NOs as N, and Puerto Rico has
statewide criteria for DIN. Delaware has criteria for DIN as N in some estuaries.
Massachusetts has not developed N criteria for estuaries but there is a TMDL in place
that limits N loading to select estuaries on Cape Cod, Martha's Vineyard, Nantucket, and
the Buzzards Bay/South Coastal area. California has made progress toward numeric
nutrient criteria. Currently, the state has numeric criteria for water clarity in all estuaries
statewide, but specific N criteria currently exist only for one location based on a TMDL
for nutrient compounds (including N) for the Malibu Creek watershed, which includes
the estuarine waters of Malibu Lagoon. By 2020, the following additional states are
expected to have at least partial numeric N criteria for estuarine waters: Connecticut,
Georgia, New York, South Carolina, and the U.S. Virgin Islands (Figure 7-10).
Currently, four states (Florida, Hawaii, North Carolina, Oregon), American Somoa and
Washington D.C. have numeric criteria for chlorophyll a, a nutrient enrichment indicator
in estuaries, and Virginia has chlorophyll a criteria for some estuaries. A summary table
June 2018
7-104
DRAFT: Do Not Cite or Quote

-------
and searchable listing of U.S. EPA-approved numeric criteria for nutrient parameters may
be found at https://www.epa.gov/nutrient-policv-data/state-progress-toward-developing-
numeric-nutrient-water-qualitv-criteria.
No N Criteria Currently Planned
Partial N Criteria
Planning to have N Criteria by 2020
Statewide N Criteria
Figure 7-10 State progress toward developing numeric nutrient criteria for
estuaries https://www.epa.aov/nutrient-policv-data/state-proaress-
toward-developina-numeric-nutrient-water-aualitv-criteria.
June 2018
7-105
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
7.2.11.
Estuary and Near-Coastal Biogeochemistry Summary
In the 2008 ISA, the evidence was sufficient to infer a causal relationship between
reactive N deposition and biogeochemical cycling of N and C in estuarine and
near-coastal marine systems. Nitrogen pollution is the major cause of harm to the
majority of estuaries in the U.S. (Brickcr et al.. 2008; NRC. 2000) and can lead to
eutrophication, the process of increasing nutrient over-enrichment leading to water
quality deterioration. Eutrophic systems are characterized by an increase in the rate of
supply of organic matter (primary production and organic carbon accumulation) in excess
of what an ecosystem is normally adapted to processing (Diaz et al.. 2013; Nixon. 1995).
Evidence reviewed in the 2008 ISA, along with new studies indicate elevated N inputs to
coastal areas can alter key processes that influence C and N cycling in near-coastal
environments. The body of evidence is sufficient to infer a causal relationship
between N deposition and the alteration of biogeochemistry in estuarine and
near-coastal marine systems which is consistent with the conclusions of the 2008 ISA.
As N fluxes to coastal areas have increased in recent decades in many parts of the U.S.,
the varying rates of different N cycling processes within estuaries themselves can also
affect the magnitude of eutrophication experienced as a result of external N enrichment.
Nitrogen additions not only cause the total pool of N to be larger, but may also perturb N
cycling in such a way that the system may exacerbate eutrophication to a greater extent
than expected based on N additions alone. Research conducted since the 2008 ISA has
shown that many of these N cycling processes are more important in the estuarine
environment than previously understood. The removal of N through denitrification is a
valuable ecosystem service in terms of constraining the extent and magnitude of
eutrophication. Additional research has established DNRA as a more important N
reduction pathway in some estuaries. Ammonium produced via DNRA can lead to
enhanced productivity and respiration, which may exacerbate hypoxia. Recent studies
indicate that DNRA rates are higher in warmer months and can also take up a larger
percentage of total N reduction activity when temperatures are higher. The roles of
sedimentary microbial processes of denitrification and anammox have been further
characterized. New research has shown that the community of N fixing microorganisms
is more diverse in estuarine and coastal waters than previously thought, and that N
fixation occurs more widely than previously assumed. Influence of benthic macrofauna
on N cycling has received increased research attention in part due to the potential for
these organisms to mitigate external N enrichment.
New studies and evidence reviewed in the 2008 ISA continue to show that at many U.S.
coastal areas, atmospheric deposition constitutes an important proportion of N inputs to
estuaries; ranging from <10 to approximately 70% of the N inputs (Table 7-9). As stated
June 2018
7-106
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
in the 2008 ISA, estuaries tend to be N limited (Elser et al.. 2007; Howarth and Marino.
2006; NRC. 2000; Nixon. 1995; Vitousek and Howarth. 1991; Howarth. 1988; D'Elia et
al.. 1986). and many currently receive high N loads from human activities
(e.g., atmospheric deposition, agricultural runoff, wastewater, and other sources), which
can cause eutrophication (Howarth et al.. 1996a; Vitousek and Howarth. 1991). It is well
known that the development and continuation of hypoxia in estuary and marine systems
can be accelerated by increased nutrient loading. Monitoring efforts across the U.S.
(Appendix 7.2.7) continue to show that N enrichment is a widespread problem. Numeric
Nutrient Criteria have been established to varying degrees by coastal states in an effort to
manage N inputs (Appendix 7.2.9). Water quality deterioration from N inputs to estuaries
including development of hypoxic zones, can be linked to biological changes. Biological
indicators of estuarine condition (e.g., chlorophyll a, HABs, macroalgae, SAV) are
described in Appendix 10.
Since the 2008 ISA, a number of papers have identified links between nutrient
enrichment and coastal acidification, and several mechanisms have been identified. One
of the initial studies found that CO2 production during decomposition of organic matter
delivered to coastal zones from rivers experiencing eutrophication has enhanced the
acidification of coastal subsurface waters in the Gulf of Mexico and the East China Sea
(Cai et al.. 2011c) and additional studies provide evidence of acidification in estuaries
due to this mechanism (Laurent et al.. 2017; Wallace et al.. 2014; Cai etal.. 2011c; Orr et
al.. 2005). In addition to microbial degradation of organic matter, respiration of living
algae and seagrasses during the night can also drive acidification. (Howarth et al.. 2014).
The CO2 produced in eutrophic estuarine waters combines with water molecules,
producing carbonic acid, which makes the water more acidic (Sunda and Cai. 2012; Cai
et al.. 2011c; Howarth et al.. 2011). Nutrient-enhanced coastal acidification tends to
occur in locations where there is either thermal or saline stratification. Acidification also
can be enhanced indirectly through the creation of anoxic waters and oxidation of HS
(Cai et al.. 2017b). Modeling of coastal acidification viaN enrichment and atmospheric
CO2 dissolution suggests that the combined effects of these two pathways are synergistic.
The body of evidence is sufficient to infer a likely causal relationship between N
deposition and increased nutrient-enhanced coastal acidification.
7-107
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
APPENDIX 8. BIOLOGICAL EFFECTS OF
FRESHWATER ACIDIFICATION
This appendix characterizes the biological effects of acidifying deposition of nitrogen (N)
and sulfur (S) in freshwater systems. Indicators of surface water chemistry are linked to
biological endpoints (Appendix 8.1) in freshwater systems experiencing either chronic or
episodic acidification (Appendix 8.2). Affected biota include plankton, invertebrates,
fish, and other organisms (Appendix 8.3). Next, documentation of biological recovery in
previously acidified systems (Appendix 8.4) is reviewed. Appendix 8J. includes levels of
deposition at which effects are manifested and empirical and modeled CLs for acidifying
deposition. A summary section with causal determinations based on a synthesis of new
information and previous evidence of biological effects of aquatic acidification is
presented in Appendix 8.6. Overall, the updated research synthesized in this ISA reflects
incremental improvements in scientific knowledge of aquatic biological effects and
indicators of acidification as compared with knowledge summarized in the 2008 ISA.
The causal relationships between acidifying deposition and biological effects on aquatic
ecosystems are now, and were in 2008, well supported.
8.1. Introduction
In the 2008 Integrated Science Assessment for Oxides of Nitrogen and Sulfur-Ecological
Criteria (2008 ISA), the body of evidence was sufficient to infer a causal relationship
between acidifying deposition and changes in freshwater biota. Changes in
biogeochemical processes and water chemistry caused by deposition of N and S to
surface waters and their watersheds (Appendix 7) have been well characterized for
several decades and have ramifications for biological functioning of freshwater
ecosystems. The 2008 ISA and studies since have shown that acidification from acid
deposition can result in the loss of acid-sensitive organisms, population declines, and
decreased species richness. This evidence is consistent and coherent across multiple
species. More species are lost with greater acidification, providing evidence of biological
gradients in effects. Both earlier and more recent studies indicate that aquatic biota in
sensitive aquatic ecosystems have been affected by acidification at virtually all trophic
levels. New information is consistent with the conclusions of the 2008 ISA that the
body of evidence is sufficient to infer a causal relationship between acidifying
deposition and changes in biota, including physiological impairment and alteration
of species richness, community composition, and biodiversity in freshwater
ecosystems.
June 2018
8-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
As reported in the 2008 ISA, effects of acidifying deposition on biotic integrity of
freshwater ecosystems can be linked to changes in several key chemical effects
indicators, including pH, dissolved inorganic aluminum (Al) concentration, calcium (Ca),
and acid neutralizing capacity (ANC). Biological effects are primarily attributable to low
pH and high inorganic Al concentration. The ANC, a measure of the overall buffering
capacity against acidification, is commonly used because it integrates overall acid status,
it is not affected by dissolved carbon dioxide (CO2), and surface water acidification
models do a better job projecting ANC than pH and inorganic Al concentrations.
However, ANC does not cause harm to biota. The usefulness of ANC lies in the
association between it and the surface water constituents that directly cause or ameliorate
acidity-related stress, in particular inorganic Al,—(Ca), and H+ (measured as pH).
Chemical factors such as pH, Ca, ANC, ionic metals and dissolved organic carbon (DOC)
are affected by acid deposition and can profoundly affect the structure and function of
biological communities in lakes and streams. Inorganic Al is minimally soluble at soil or
surface water pH above about 6.0, but solubility increases markedly as pH values drop
below about 5.5. Low concentrations of base cations like Ca enable low pH and high
concentrations of inorganic Al to occur (Baker etal.. 1990bV However, DOC can bind to
inorganic Al ions and reduce their bioavailability and their toxic impact on aquatic biota.
The base cation surplus (BCS) is an alternate index that is similar to ANC and that also
adjusts for the organic acid status of surface waters (Lawrence et al.. 2007). BCS is based
on a measurement of ANC (calculated from the charge balance of ionic concentrations in
water) and also accounts for the influence of natural organic acidity by including the
strongly acidic organic acid anions in the calculation.
Acid-sensitive freshwater systems can either be chronically acidified or subject to
occasional episodes of decreased pH, decreased ANC, and increased inorganic Al
concentration (Appendix 7). Changes to flow and surface water chemistry, characteristic
of these events, reflect the influence of acidic inputs from precipitation, gases, and
particles, as well as local geology and soil conditions. As stated in the 2008 ISA, surface
water chemistry is a good indicator of the effects of acidification on the biotic integrity of
freshwater ecosystems because it integrates the sum of soil and water processes that
occur within a watershed. Surface water chemistry reflects and integrates N saturation,
forest decline, soil acidification, nutrient cycling, and land use (U.S. EPA. 2003).
Acidification studies reviewed in the 2008 ISA included laboratory experiments,
bioassays, mesocosm exposures, field observations, and whole ecosystem acidification or
deacidification studies. Baker et al. (1990a) conducted a rigorous review of the effects of
acidification on aquatic biota for the 1990 National Acid Precipitation Assessment
Program (NAPAP) State of Science/Technology report series. In the Baker et al. (1990a)
8-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
report, hundreds of laboratory studies, in situ bioassays, field surveys, whole-system field
experiments, and mesocosm studies on the effects of acidification on aquatic biota were
evaluated. The findings in that report, along with literature published after 1990 up to
December 2007, were assessed in the 2008 ISA. Measures of health, vigor, reproductive
success, and biodiversity of aquatic biota were identified in the 2008 ISA as being
affected by acidified waters caused by acidifying deposition. Effects had been most
clearly documented for fish, aquatic invertebrates, and algae.
Since publication of the 2008 ISA, the overarching understanding of aquatic acidification
has not changed appreciably. More recent research has confirmed and strengthened this
understanding and provided more quantitative information, especially across the regional
landscape. This appendix highlights post-2007 research findings through the spring of
2017.
8.2. Chronic versus Episodic Acidification
Traditionally acidification involves both chronic and episodic processes. As defined in
the 2008 ISA, chronic acidification refers to annual average conditions, which are often
represented as summer and fall chemistry for lakes and as spring baseflow chemistry for
streams. Episodic acidification refers to conditions during rainstorms or snowmelt when
proportionately more drainage water is routed through upper soil horizons, which tend to
provide less neutralization of atmospheric acidity as compared with deeper soil horizons.
Surface water chemistry exhibits lower pH and ANC during episodic than during
baseflow conditions. Chronically acidic lakes and streams maintain ANC <0 |icq/L. on
average, throughout the year (Driscoll et al.. 2001b). They are no longer prevalent in
regions of the U.S. affected by acidic depostion (cfi, Fakhraei et al.. 2016; Fakhraei et al..
2014). The ANC during acidic episodic events may fall below 0 |icq/L for only a few
hours to weeks in a given year. It is known that the biota in many streams/lakes are
impacted when the ANC is consistently below 50 ueq/L. For this reason, the U.S. EPA
National Lakes Assessment used an ANC threshold of >50 ueq/L as indicative of
nonacidified water bodies (U.S. EPA. 2009b). In addition, dynamic models, such as
MAGIC and PnET-BGC, estimate that some lakes and streams in the Adirondack and
southern Appalachian Mountains had preindustrial ANC below 50 (ieq/L, but few or
none had preindustrial ANC below 20 (ieq/L.
A large portion of the available aquatic acidification data reported in the 2008 ISA
focused on lakes and streams in the Northeast and the southern-Appalachian Mountains
(U.S. EPA. 2008a') because these systems have been among the most impacted by
acidifying deposition in the past and have the best available surface water monitoring
June 2018
8-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
information. Since completion of the literature review for the 2008 ISA, additional
research has been conducted on changes in chronic surface water chemistry in the U.S. in
response to changing levels of acidic deposition (Appendix 7). Many of these studies
have further documented chemical recovery as S and N deposition continue to decline in
most areas in the U.S. Biological recovery has been observed in some systems, but
available data generally indicate that it typically lags behind chemical recovery
(Appendix 8.4).
Many streams that exhibit chemical conditions during base flow that are suitable for
aquatic life are subject to occasional episodic acidification that may exceed the acid
tolerance of many aquatic species. Episodic events vary in timing, duration, and intensity.
Episodic acidification is most common in the early spring and late fall, and least common
in summer, when high flows tend to be infrequent (Driscoll et al.. 2001b). This
seasonality of episodic events reflects to some extent the influence of deposition
accumulating in the landscape that is flushed to drainage water during precipitation or
snowmelt events. Other natural processes can also be involved, including altered
hydrologic flowpaths, biological uptake, and the neutral salt effect, whereby deposition of
a neutral salt (e.g., NaCl) can lead to ion exchange of H+ for Na+ in soil, followed by
drainage water acidification. Episodic processes are mostly natural, but SOr and NOs
influxes due to atmospheric deposition play important roles in the episodic acidification
of some surface waters. Dilution of base cation concentrations in runoff during
episodes—a quantitatively important component of the episodic response—may be
affected in part by past base cation depletion of watershed soils, due to acidification,
which limits the release of Ca and Mg in response to rapid runoff during storms and/or
snowmelt (Wigington et al.. 1996a; Wigington et al.. 1996b).
As reported in the 2008 ISA, episodes are generally accompanied by changes in two or
more of the following chemical parameters: ANC, pH, concentrations of base cations,
SO42 , NO3 . aluminum ions, organic acid anions, and DOC (Sullivan. 2000). During
short-term episodes (hours or days), both flow and water chemistry can change markedly.
Biological effects of episodes sometimes include fish mortalities, changes in species
composition, and declines in aquatic species richness across multiple taxa, ecosystems,
and regions. The U.S. EPA's Episodic Response Project (ERP) in the 1980s confirmed
the chemical and biological effects of episodic pH depressions in lakes and streams in
parts of the U.S. (Wigington et al.. 1993). In the ERP report, streams having acidic
episodes showed long-term effects on fish populations compared with streams in which
ANC remained above 0 (j,eq/L. Results reported in the 2008 ISA from in situ bioassay
studies conducted across the eastern U.S. showed that acidic episodes (with associated
low pH and elevated inorganic Al concentrations and high stream-water discharge)
caused rapid fish mortality under some conditions (Driscoll et al.. 2001b; Bulger et al..
8-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
1999; Baker etal.. 1996). Baker et al. (1990a) concluded that episodes are most likely to
affect biota if the episode occurs in waters with pre-episode pH above 5.5 and minimum
pH during the episode of less than 5.0. In a later study, acid episodes reduced the size of
fish populations and eliminated acid-sensitive species if median high-flow pH was less
than 5.2 and inorganic Al concentration exceeded 100 (ig/L (3.7 (j,M), despite the
relatively short duration of the episodes studied (Baker etal.. 1996). Research from
several regions in the U.S. indicated that acidifying deposition likely has increased the
magnitude, frequency, and biological effects of episodic acidification events.
8.3. Aquatic Organisms Impacted by Acidifying Deposition
Appendix 8.3.1 through Appendix 8.3.8 describe effects of acidification on
phytoplankton, zooplankton, benthic invertebrates, fish, birds, and other biota. Changes
in biota are linked to chemical indicators (Appendix 7) in surface water. In the 2008 ISA,
biological effects were divided into two major categories: (1) effects on health, vigor, and
reproductive success of taxonomic groups and (2) effects on biodiversity. The first
category included changes at the species level of biological organization such as
cumulative sublethal physiological effects (individual condition factor) and recruitment
success. Effects on biodiversity include changes in community structure, species
composition, and taxonomic richness. Studies reviewed in the 2008 ISA showed that the
earlier aquatic lifestages were particularly sensitive to acidification.
In Appendix 8.3.1 through Appendix 8.3.8. findings from the 2008 ISA are summarized
for each major taxonomic group, followed by a review of new literature including studies
that report physiological alterations, changes in the abundance or presence of taxa,
community shifts, and other biological responses associated with acidifying conditions.
Effects on organisms at lower trophic levels, such as plankton and invertebrates, are
discussed first followed by observations for fish and other vertebrates. Ecological
thresholds of chemical indicator(s) associated with the observed responses are included,
when available, from the reviewed studies. Appendix 84 considers the evidence for
biological recovery of different taxa.
8.3.1. Plankton
Plankton, floating or drifting organisms in the water column, play an important role in
freshwater ecosystems. Phytoplankton, or suspended algae, are primary producers at the
base of the aquatic food web. These photosynthetic organisms, encompassing diatoms,
cyanobacteria, dinoflagellates, and other groups of algae, vary in tolerance of acidic
June 2018
8-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
conditions. Zooplankton comprise many groups of freshwater organisms including
protozoans, rotifers, cladocerans, and copepods. Zooplankton feed on phytoplankton or
other zooplankton. Abundance and community composition of plankton respond to
changes in surface water chemistry associated with acidifying deposition to water bodies.
8.3.1.1. Phytoplankton
Studies reviewed in the 2008 ISA reported reduced species richness of phytoplankton
with decreases in pH and increases in inorganic A1 concentrations associated with
acid-affected surface waters (U.S. EPA. 2008a'). There was also a shift in the composition
of dominant taxa, but species composition shifts could not be accurately predicted. This
kind of effect was most prevalent in the pH 5 to 6 range (Baker etal.. 1990a). It appeared
that the phytoplankton community restructured as the water became acidified. However,
the response of phytoplankton communities often vary, with some lakes showing
increasing biomass, others decreasing, and others having no change in phytoplankton
biomass (Baker et al.. 1990a). Leavitt et al. (1999) suggested that the complex
interactions between pH, DOC, and light can explain the high variability in
phytoplankton-biomass-acidification relationships.
More recent studies of the responses of phytoplankton to changes in surface water acidity
have been limited. Effects on primary productivity are uncertain. In regards to
phytoplankton species composition, changes were not apparent in the lake diatom
communities or the diatom-inferred lake pH of sequential core segments in four
alpine/subalpine lakes in the Cascade Mountains of Washington and Oregon that have not
experienced substantial water acidification (Eilers et al.. 2016). The study lakes were
very low in specific conductance (<3.6 (iS/cm) and ANC (<11 j^icq/L). Sediment cores
were collected from each lake and analyzed for nutrients and diatom microfossils in dated
(210-Pb and 14-C) core segments. Water chemistry was simulated using a modified
version of the CE-QUAL-W2 model to account for the assumed large influence of in-lake
chemical processes. Model projections of future chemistry suggested that the three study
lakes in Oregon will not change their acid-base chemistry in the future under N and S
deposition increases to 3 times ambient, largely due to long lake water residence times,
which allow in-lake processes to neutralize acid inputs. Foehn Lake in Washington was
much more acid-sensitive, increasing in acidity in response to projected increases in N
and S deposition at levels of 50% above ambient. Shallow depth, large amount of
exposed bedrock, and sparse sediment accumulation enhance its acid sensitivity. This
lake was formed less than a century ago, probably in response to climate warming and
melting of an adjacent snowfield.
June 2018
8-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Lacoul etal. (2011). reviewed information on the effects of acidification on plankton and
other organisms in Atlantic Canada and observed that the greatest changes in
phytoplankton species richness occurred over a pH range of 4.7 to 5.6, just beyond the
interval (pH 5.5 to 6.5) where bicarbonate becomes rapidly depleted in the water. Under
acidifying conditions, phytoplankton communities shifted from dominance by
chrysophytes, other flagellates, and diatoms to dominance by larger dinoflagellates.
However, biomass and productivity were not much affected. Algal biomass in five
Pennsylvania streams decreased with the severity of episodic acidification (Macdougall ct
al.. 2008). Although no significant trends were observed for taxon richness or diversity,
the degree of acidification during episodes appeared to affect the quantity of algae in the
streams. Bloch and Wevhenmever (2012) evaluated physical and chemical time-series
data for 13 nutrient-poor Swedish lakes across a latitudinal gradient. Phytoplankton
biomass and species richness showed only weak relations with water chemistry. The only
significant association observed across seven of the surveyed lakes was the number of
chlorophyte taxa and changes in water temperature and color.
As stated in the 2008 ISA, diatoms, which comprise an important component of the
phytoplankton, are excellent biological indicators of environmental change in aquatic
ecosystems, being sensitive to changes in acidity, nutrient status, salinity, and climate
(Stoermer and Smol. 1999; Sullivan and Charles. 1994). Since the 2008 ISA, studies have
further linked phytoplankton community shifts to acidification using paleolimnological
analysis of fossil diatoms or long-term sampling of phytoplankton from the water
column. Most of these studies have documented phytoplankton recovery in historically
acidified lakes (Appendix 8.4.1). Stratigraphy of sediment chrysophyte remains in
Brooktrout Lake in the Adirondack Mountains, NY revealed shifts in Mallomonas spp.
and Synura spp., with some species declining and others increasing. After the 1950s,
Fragilariforma acidobiontica, a diatom that is often abundant at pH <5.0, was present in
lake sediments. These observations in Brooktrout Lake correspond to historical
acidification, which was most severe during the 1970s and 1980s (Sutherland ct al..
2015). Phytoplankton have also been used as indicators of acidification and recovery in
Canada and Europe. In a recent analysis of diatom fossils from 10 lakes located about
80-250 km east and northeast of the oil sands development near Fort McMurray and Fort
Mackay, northwestern Saskatchewan, Laird et al. (2013) observed increases in scaled
chrysophytes and diatom flux rates in post-1980 sediments potentially related to
atmospheric deposition in the oil sands region. A slight decrease (0.25 pH unit) in
diatom-inferred pH occurred in one study site closest to the oil sands development, but
there was no evidence of widespread acidification. Overtime, phytoplankton
assemblages in Swedish lakes recovering from acidification have become more similar to
those in reference lakes (Johnson and Angeler. 2010).
8-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
8.3.1.2. Zooplankton
In studies reviewed in the 2008 ISA, decreases in ANC and pH and increases in inorganic
A1 concentration were shown to contribute to the loss of zooplankton species or
decreased abundance in lakes (Keller and Gunn. 1995; Schindler et al.. 1985). A decrease
in pH from 6 to 5 reduced species richness in lake zooplankton communities (Holt et al..
2003; Holt and Yan. 2003; Locke and Sprules. 1994). Sullivan et al. (2006a) found that
zooplankton communities varied with ANC in Adirondack lakes, with lower taxonomic
richness (number of species of crustaceans, rotifers, and total zooplankton) in lakes
having lower ANC. In general, lake-water ANC explained nearly half of the variation in
total zooplankton and crustacean taxonomic richness, but less for rotifer richness.
Particularly low zooplankton community richness occurred when ANC levels were below
0 |icq/L (15 species in highly acidic lakes compared to 35 at the highest values of ANC in
the study [near 200 |icq/L|). Observations from in situ enclosure acidification studies at
Emerald Lake in the Sierra Nevada showed shifts in zooplankton community with
decreased pH (Barmuta et al.. 1990). Daphnia rosea and Diaptomus signicauda were
eliminated below pH 5.0 while other species such as Bosmina longirostris and Keratella
taurocephala became more abundant. Possible mechanisms for zooplankton sensitivity to
low pH and ANC include ion regulation failure, reduced oxygen uptake, inability to
reproduce, and Al toxicity (U.S. EPA. 2008a).
A number of studies have been conducted since the 2008 ISA to determine the response
of zooplankton to lake acidification. Highlighted here are several studies conducted in the
U.S. and Canada. Many of these studies indicate that multiple factors could influence
zooplankton community changes. Bosmina is among the most common North American
temperate lake pelagic invertebrate genera. It is a genus of cladoceran filter feeders that
can change body size and appendage length over multiple generations in response to
changing environmental conditions. This makes it a good candidate ecological indicator.
Labai et al. (2016) evaluated Bosmina size responses in lakes near Sudbury, Ontario,
Canada, that had been acidified and then chemically recovered in response to changes in
nearby metal smelter emissions. Even with the recent return to presmelter lake pH, the
Bosmina size structure has not yet recovered to preacidification conditions. Labai et al.
(2016) suggested that the observed effects of acidification and deacidification on the size
of Bosmina may have been mediated by food web dominance of small copepod predators.
Vinebrooke etal. (2009) reported variations in zooplankton communities during a
whole-lake experimental acidification of Lake 302S in the Experimental Lakes area in
Ontario, Canada. There was a negative effect on zooplankton species richness as pH
decreased from 6.8 to 4.5. However, no correlation was found between functional
properties (productivity or net total biomass) and species richness with this pH change,
indicating that other factors such as multiple stressor interactions, species occurrences,
June 2018
8-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
and altered trophic interactions might also influence zooplankton community change. In
northwestern boreal shield lakes downwind of N emissions from the Athabasca oil sands
extraction in Saskatchewan, Canada, zooplankton community structure was influenced by
local environmental factors, including acidity, predation, and lake productivity. The role
of regional atmospheric deposition on zooplankton community composition could not be
distinguished from natural variability due to a lack of baseline data prior to oil extraction
operations (Anas et al.. 2014). Thus, the role of acidic deposition in driving changes in
zooplankton community composition in that region was not clear.
Acidification often reduces Ca availability in lake water, which may impact invertebrates
that require Ca for growth such as Daphnia spp. and crayfish (Appendix 8.4.3). Jeziorski
et al. (2012b) examined the growth and survival of daphnid species across a Ca gradient
(from 50 to 150 j^icq/L) in central Ontario soft-water lakes. Considerable variability in
growth and survival was observed within the Daphnia pulex species complex, and the
variation across all cladoceran genera was best explained by pH and lake depth. This
research adds support to the previous observation that among the cladocerans, daphnids
are especially sensitive to decreases in Ca which leads to declines in their growth and
survival as lake water acidifies with inputs of acidifying deposition.
8.3.2. Periphyton
Periphyton mats are biofilms of algae, cyanobacteria, fungi, microinvertebrates, organic
detritus, inorganic particles, and heterotrophic microbes imbedded within a matrix and
attached to submerged substrates in aquatic systems (e.g., stream or lake bottoms or
submerged vegetation). Periphyton is an important food source for invertebrates,
tadpoles, and some fish. As reported in the 2008 ISA, acidification impacts periphyton
species differently, with some excluded from impacted water bodies while others become
dominant. Such changes decrease species richness and alter community structure (U.S.
EPA. 2008a). For example, many of the brown algae and cyanobacterial periphyton
species cannot tolerate acidic conditions, causing their abundance to decline along with
pH, while green algae, particularly the filamentous Zygnemataceae, increase in relative
abundance at lower pH (Baker etal.. 1990a). There is evidence that the biomass of
attached periphyton increases at lower pH. No new studies of acidifying effects on
periphyton were identified for this review.
June 2018
8-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
8.3.3.
Benthic Invertebrates
Benthic invertebrates live in freshwater sediments and include groups such as bivalves,
worms, gastropods and insect larvae. As reviewed in the 2008 ISA, acidification as
measured by decreases in ANC and pH and increases in inorganic A1 concentration has
been shown to contribute to the loss or decline in abundance of benthic invertebrate
species in streams. The Ephemeroptera (mayflies)-Plecoptera (stoneflies)-Tricoptera
(caddisflies) (EPT) Index is a common measure of stream macroinvertebrate community
integrity. The EPT metric is calculated as the total number of families present in those
three insect orders. Mayflies tend to be the most sensitive of the three, and stoneflies tend
to be the least sensitive (Peterson and Van Eeckhaute. 1992). The mayfly order alone is
often selected for study because it includes a number of genera and species having
varying degrees of sensitivity to acidification, including some that are highly sensitive
(Sullivan et al.. 2003). Typically, pH values below 5 result in the virtual elimination of all
mayflies along with some other aquatic organisms from some streams (U.S. EPA. 2008a;
Baker and Christensen. 1991). These benthic invertebrates are impacted by acidification
because H+ and inorganic Al can be directly toxic, causing disruption of their ion
regulation and reproductive success. U.S. EPA first published freshwater criteria for Al in
1988 and is in the process of updating the Aquatic Life Criteria. The current 1988 acute
and chronic criteria are 750 and 87 (ig/L, respectably, not be exceeded once every 3 years
on average (U.S. EPA. 1988).
Work since the 2008 ISA by Baldigo et al. (2009) assessed effects of acidification on
benthic macroinvertebrate community dynamics in the southwestern Adirondack
Mountains. Water chemistry and benthic macroinvertebrates were surveyed in 36 streams
with different levels of acidity to characterize community response and identify
thresholds for biological effects (Figure 8-1). In the study streams, macroinvertebrate
assemblages were severely impacted at pH <5.1, moderately impacted at pH from 5.1 to
5.7, slightly impacted at pH from 5.7 to 6.4 and usually unaffected above pH 6.4.
Inorganic Al concentrations reached potentially toxic levels in two-thirds of the study
streams. The authors applied the Acid Biological Assessment Profile [acidBAP] index,
which is based on percentage mayfly richness and percentage acid-tolerant
macroinvertebrate taxa (Burns et al.. 2008b). The acidBAP was strongly correlated with
pH, ANC, BCS, and the concentration of inorganic Al, indicating the loss of mayflies in
the stream as the surface waters pH declines from pH 7.0 to 4.2. A loss of about
12 species occurred in streams that had a pH between 7 and 4.2. Regression across all
36 streams showed a loss of 4.6 species per unit pH decrease (Figure 8-1). Inorganic Al
toxicity was likely the main cause of the loss of macroinvertebrates. The Al concentration
is strongly correlated with surface water pH (as pH decreases solubility of inorganic Al
increases) and acid-base balanced as measured by BCS.
June 2018
8-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Several pH thresholds for aquatic invertebrate response have been published since the
2008 ISA (Table S-1). Lacoul et al. (2011) reviewed available information on the effects
of water acidification on aquatic organisms in Atlantic Canada. The median pH for
sensitive invertebrate species occurrence was between 5.2 and 6.1, below which species
tended to be absent. For example, several species of mayfly and most gastropods are
intolerant of acidity and only occur at pH >5.5 and >6, respectively.
The make-up of the stream invertebrate community is governed in part by the condition
of riparian vegetation and associated humic acid levels in surface water. O'Toole et al.
(2017) investigated riparian vegetation condition under high and low humic acid
influence as affected by riparian vegetation. Both perennial and intermittent streams in
western Australia were evaluated. Streams having well-developed riparian vegetation
showed proportionately more algal grazers and detritivores. Intermittent streams with
high humic content had lower numbers of cladocerans and chironomids and higher
numbers of grazing gastropods as compared with intermittent streams that had relatively
low humic content.
Several European studies published since the 2008 ISA have evaluated the use of benthic
invertebrates as biological metrics to classify the ecological status of water bodies. These
studies have developed or applied indices to predict dose-response relationships between
macroinvertebrate communities and water chemistry. One study by Schartau et al. (2008)
applied existing macroinvertebrate metrics developed for river acidification to lakes and
also developed and tested new species-based indicators of lake acidification based on
668 samples of littoral macroinvertebrates taken from 427 lakes across Sweden, the U.K.,
and Norway. Although there was high variation in the data, a response threshold between
pH 5.8 and 6.5 was identified for littoral macroinvertebrate communities, suggesting they
could be used to assess ecological quality of lakes.
8-11
DRAFT: Do Not Cite or Quote

-------
40
35 -|
30
25
Ł 20
{l)
9- 15
TO
o 10
F
5
0


•


s •
* *
non-impacted
• • •
•
•
i" * *
slight impact •
t #
• «
•

#
•
•
•


moderate impact


severe impact

y = 4.62x - 1.49
R2 = 0.57
4.0
4.5
5.0
5.5	6.0
Median pH
6.5
7.0
7.5
Source: Modified from Baldiao et al. (2009)
Figure 8-1
Total macroinvertebrate species (community) richness as a
function of median pH in 36 streams sampled in the western
Adirondack Mountains of New York, 2003-2005; the four standard
(New York State) impact categories for species richness are
defined.
June 2018
8-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
Table 8-1 Thresholds of biological response to changes in water acidity for
benthic invertebrates published since the 2008 Integrated Science
Assessment for Oxides of Nitrogen and Sulfur-Ecological Criteria.
Life Forms
Region
Potential
Thresholds
Reference
Sensitive invertebrates Atlantic Canada
present
pH 5.2 to 6.1 Lacoul et al. (2011)
Littoral macroinvertebrates Northern Europe
pH 5.8 to 6.5 Schartau etal. (2008)
Stream macroinvertebrates Southwestern Adirondacks pH 5.1 to 5.7 Baldiao et al. (2009)
Mountains
Stream macroinvertebrates Sweden
pH 5.7 to 6.0 Andren and Wiklund (2013)
Using macroinvertebrate monitoring data (1,462 samples) from Northern Europe Moe et
al. (2010) tested new and existing acidification metrics with the aim of selecting a
common metric to assess pH effects on biota in this region. Most investigated metrics
responded to pH, with most of the variance explained by a few variables, including the
number of Ephemeroptera families and the proportion of sensitive Ephemeroptera. Most
biological metrics were higher (less impacted) in humic than in clear waters, suggesting
smaller acidification effects in humic waters. Murphy et al. (2013) described an approach
for developing diagnostic indices for assessing acidity in British streams that could be
useful in the U.S. Variation in macroinvertebrate assemblages in 76 test sites was
quantified with a 197-site calibration set. The Acid Water Indicator Community Index
expressed at the species level was related to base-flow pH and to storm-flow pH and
ANC, accounting for 38 to 56% of the variation in acid condition.
Interactions among Al, pH, and organic matter, and their collective influence on toxicity
to aquatic biota, are discussed in Appendix 8.3.6. In general, decrease in pH brings larger
amounts of Al into solution whereas when dissolved organic matter is also present, it
binds with Al, converting it into organo-Al complexes, decreasing its toxicity. In six
humic (mean total organic carbon [TOC] 10-18 mg/L) streams in central Sweden,
Andren and Wiklund (2013) evaluated acute and sublethal toxicity of macroinvertebrate
salmonid prey items exposed to episodes of acidity during spring snowmelt. The
threshold for mortality was pH <5.7 and inorganic Al >20 (ig/L (0.7 (iM) for mayflies
(Baetis rhodani) and pH <6.0 and inorganic Al >15 (ig/L (0.6 (j,M) for the freshwater
amphipod Gammarus pulex. Traister et al. (2013) investigated shifts in macroinvertebrate
communities and food webs in nine small forested streams in the Czech Republic across a
June 2018
8-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
pH gradient from 4.0 to 7.7. Acidification was related to reduced taxon richness and
reduced Ephemeroptera family richness and lower population densities.
8.3.4. Bacteria, Macrophytes, and Bryophytes
At the time of the 2008 ISA, most observations of biological effects of acidification were
for phytoplankton (Appendix 8.3.1.1). benthic invertebrates (Appendix 8.3.3). and fish
(Appendix 8.3.6). and relatively little information was available regarding the response of
other biological taxato surface water acidification. This has not changed appreciably. A
few studies have been conducted on bacteria, macrophytes (aquatic plants), and
bryophytes, as described below.
Percent et al. (2008) assessed bacterioplankton community diversity and structure in
18 Adirondack lakes across a range of acid-base chemistry using sequencing and
amplified ribosomal DNA restriction analysis of constructed rRNA gene libraries. Based
on principal components analysis, pH was positively correlated with bacterioplankton
community richness and diversity. The richness of several bacterial classes, including
Alphaproteobacteria, was directly correlated with pH. However, other environmental
factors, such as lake depth, hydraulic retention time, dissolved inorganic C, and nonlabile
(organically bound) monomeric Al, were also important in explaining bacterioplankton
community richness and diversity, suggesting that acidity is only one factor that controls
community composition.
Macrophyte studies provide additional information on the biological responses to acidic
conditions in lakes. For example, a pH threshold for the absence of sensitive macrophytes
was determined to be 5.5 in Atlantic Canada lakes by Lacoul etal. (2011). In an analysis
of factors affecting the distribution of aquatic macrophytes in lakes in the Pyrenees
Mountains of southern Europe, Pulido et al. (2015) constructed statistical models to
predict optimum ranges and ecological niches of 11 aquatic macrophyte species.
Macrophytes were most suited to lakes with low water concentrations of NOs and SOr
in vegetated watersheds at low elevation. It was possible to delineate individual species
ranges along gradients of specific conductance and water pH.
Bryophytes, most notably mosses and liverworts, are nonvascular plants that often
constitute an important component of the biodiversity and productivity of low-order
streams. Tessler etal. (2014) assessed bryophyte assemblages in southeastern New York
streams and found that some species like Hygrohypnum eygyrium and Codriophorus
aduncoides were generalists and able to tolerate pH in the range from approximately 4 to
7, whereas others were more strongly limited by pH. For example, H. ochraceum
occurred only at circumneutral pH near 6.5, while Andreaea rothii was restricted to pH
June 2018
8-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
<5. Stream pH and amount of bedrock substrate were identified in this study as the
primary determinates of byrophyte assemblage composition.
8.3.5. Amphibians
Amphibians such as frogs, newts, toads, and salamanders have aquatic lifestages and may
be in contact with acidified waters in areas affected by acidifying deposition. Although
some species of amphibian are on the decline, in the 2008 ISA, there was no evidence to
suggest that acidic deposition was an important factor that impacts the health and
abundance of amphibian communities (U.S. EPA. 2008a'). Nevertheless, there are both
relatively acid-sensitive and acid-tolerant amphibians. Examples of acid-sensitive
amphibians in Baker etal. (1990a) include the spotted salamander (Ambystoma
maculatum) and Jefferson salamander (Ambystoma jeffersonianum). Jefferson
salamanders were absent from ponds with very low pH (<4.5; Freda and Dunson. 1986).
Based on transplant studies into ponds or to the laboratory, embryos of this species did
not hatch in water with pH less than about 4.5. Acid-tolerant embryos such as the Pine
Barrens treefrog (Hyla andersonii) may hatch at a pH of 3.7 (Freda and Dunson. 1986).
Toxicity is not solely a matter of pH, but is also influenced by Ca2+, inorganic Al, and
DOC concentrations, lifestage, and water temperature (Baker etal.. 1990a). Large-scale
amphibian extinctions due to acidifying deposition had not been detected in any
geographic region at the time of the 2008 ISA (Baker etal.. 1990a) and have not been
documented in more recent years.
Studies published since the 2008 ISA further indicate that amphibian species are
relatively tolerant of acidifying conditions. In a review of toxicity data for amphibian
species found in Atlantic Canada, Lacoul et al. (2011) concluded that some amphibians
can survive at pH as low as 3.5 to 4.0. Chambers et al. (2013) explored the relationship
between pH and hormonal response in salamanders. In their field studies with Jefferson
salamander larvae from eight natural breeding populations and a mesocosm experiment
with varying pH, they observed an increase in baseline corticosterone concentration with
lower pH (in the range of 5 to 5.8). Elevation of baseline corticosterone level is a
physiological alteration associated with stress response. In contrast, no significant
relationship was observed between corticosterone and pH in adult Allegheny Mountain
dusky salamanders (Desmognathus ochrophaeus) living in nine streams in the Central
Appalachian ecoregion with differing pH (Woodlev et al.. 2014). Three-week laboratory
exposures with the same species showed that low pH (3.5) decreased locomotory activity
but had no effect on plasma corticosterone levels. A change of average pH from pH 7 to
pH 6 resulted in a shift in the skin microbial community composition on larval American
bullfrogs (Rana catesbeiana) (Krvnak et al.. 2015). Following metamorphosis, shifts in
June 2018
8-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
pH did not alter skin microbial community structure significantly in juvenile frogs.
However, antimicrobial peptide production was affected by interactions between pH and
degree of shading, suggesting that environmental variability may influence amphibian
susceptibility to fungal pathogens.
8.3.6. Fish
Physiological and population-level responses associated with exposure of fish to acidified
waters have been well characterized for many decades. As summarized in the NAPAP
report by Baker et al. (1990a) and studies reviewed in the 2008 ISA, fish populations in
acidified streams and lakes of Europe and North America have declined, and some have
been eliminated due to atmospheric deposition of acids and the resulting decrease in pH
and ANC and increase in inorganic Al concentrations in surface waters. By 1990, it was
well established that freshwater acidification could cause significant adverse biological
effects on fish, although the effects were not uniform across species. Summary
information from the 2008 ISA for pH, ANC, and Al and effects on fish are presented in
Appendix 8.3.6.1 to Appendix 8.3.6.5 along with results of new studies. In general,
understanding of the effects of acidification on fish has not changed since the 2008 ISA.
Effects of acidification on fish and other organisms must be viewed in the context of
other stressors and management actions, in particular climate change and the effects of
fish stocking.
Responses among fish species and lifestages within species to changes in pH and Al in
surface waters are variable. In general, early lifestages such as larvae and smolts are more
sensitive to acidic conditions than the young-of-the-year, yearlings, and adults (Baker et
al.. 1990a; Johnson et al.. 1987; Baker and Schofield. 1985). Some of the most commonly
studied species have been brown trout (Salmo trutta), brook trout (Salvelinus fontinalis),
and Atlantic salmon (Salmo salar). Among these three species, Giedrem and Rosseland
(2012) recently showed substantial additive genetic variation in tolerance to acidic water,
with heritabilities (h[2]) ranging from 0.09 to 0.27 for dead-eyed eggs (a development
period that is highly sensitive to low pH [4.7 to 5.2]).
Several new studies of Atlantic salmon provide additional information documenting the
sensitivity of this species to acidification during different lifestages. Evidence suggests
that acidification has been an important stressor that has limited the distribution and
abundance of Atlantic salmon in the northeastern U.S. Atlantic salmon are anadromous
and they are susceptible to physiological effects of acidification as they develop and
move to different habitats. The salmon start their life cycle in freshwater then migrate to
the ocean, returning to freshwater to spawn. In rivers, eggs hatch into fry, which develop
June 2018
8-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
into parr. Before migrating to the ocean, parr start developing into smolt, which are more
sensitive to acidification than the parr lifestage (Kroglund et al.. 2008; Monette and
McCormick. 2008). Recent research has shown that exposure to concentrations of
inorganic Al that have no apparent effects in freshwater may subsequently affect smolt
survival in seawater. Thus, the timing of acidification episodes in relation to fish lifestage
and migration from freshwater to seawater may impact fish survival due to the delayed
response to inorganic Al exposure (Kroglund et al.. 2008; Kroglund et al.. 2007).
8.3.6.1. Physiological Responses to Acidification
The modes of action of biological impacts on fish from surface water acidification were
reasonably well known at the time of the 2008 ISA. Physiological disturbances to fish
exposed to acidic waters reported in the 2008 ISA included iono- and osmoregulatory
failure, acid-base regulatory failure, and respiratory and circulatory failure. These
impacts can often be directly attributed to effects on gill function or structure. Al has
been shown to accumulate on the gill surface when fish are exposed to water having high
inorganic Al concentration. The primary mechanism for the toxic effects of low pH and
elevated inorganic Al on fish involves disruption of normal ion regulation at the gill
surface, resulting in increased rates of ion loss and inhibition of ion uptake (Bergman et
al.. 1988; Wood and McDonald. 1987; Leivestad. 1982; McWilliams and Potts. 1978).
The disruption of salt and water balance causes red blood cells to rupture and blood
viscosity to increase (Driscoll et al.. 2003b). Additional effects might include disruption
of Ca metabolism (Reader et al.. 1988; Gunn and Noakes. 1987; Peterson and Martin-
Robichaud. 1986) and loss of Ca from important binding sites in the gill epithelium,
which reduces the ability of the gill to control membrane permeability (Exlcv and
Phillips. 1988; Havas. 1986; Mcdonald. 1983).
Newly available literature since the last review supports previous findings of effects of
acidification on fish physiology. Several in situ studies on brook trout in the southeastern
U.S. and New England provide additional information on sensitivity of different fish
lifestages to episodic acidification. Changes in native brook trout physiology were
determined during two acid runoff episodes in the Great Smoky Mountains National Park
by Neff et al. (2008). Results of an in situ bioassay of whole-body sodium concentrations
before and after acidification showed that stream acidification negatively impacted native
trout physiology. Loss of whole-body sodium when stream pH dropped below 5.1
indicated that trout lost the ability to ionoregulate. Stream water ANC and pH decreased
episodically during the stormflow events studied. An ANC contribution analysis
indicated that acidic deposition may have been the major cause of episodic stream
acidification; increased concentrations of organic acid anions and base cation dilution
June 2018
8-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
during episodes also appeared to be important. Subsequently, Neff et al. (2009)
investigated the effects of hydrologic episodes on loss of blood plasma Na+ in native
southern brook trout, a response associated with physiological stress from acid exposure.
In situ bioassays were conducted at three sites during a 2-year period. Whole-body Na+
concentrations decreased by 10 to 20% following acidic episodes during which 24-hour
mean pH values of 4.88, 5.09, and 4.87 and total dissolved Al concentration of 210, 202,
and 202 (ig/L (7.8, 7.5, and 7.5 (iM) were observed.
In a combination lab and field study designed to establish whether the smolt or parr
lifestage is more sensitive to short-term acid/Al pulses, Monette and McCormick (2008)
observed that compared with control fish, smolts exposed to elevated acid/Al levels
showed greater losses of blood plasma Cl~ (9-14 mM) after 2 and 6 days and increases in
plasma Cortisol (4.3-fold) and glucose (2.9-fold) levels after 6 days of exposure, while
parr were not affected. Gill Na+/K+-ATPase (NKA) activity was not affected in either
lifestage. Smolts were shown to be more sensitive than parr to short-term acid/Al
exposure although gill accumulation of Al was observed in both lifestages.
McCormick et al. (2009) determined the effects of pH and Al on survival, development
of smolts, ion regulation, and stress levels of Atlantic salmon in southern Vermont. Two
6-day field studies during the peak of smolt development (late April and early May) were
conducted in five streams having different acid-base chemistry. The researchers found
increased mortality, loss of blood plasma chloride (Cr), altered NKA activity in the gills,
and higher gill Al in fish that were caged in streams that experienced low pH (5.4-5.6)
and high inorganic Al concentration (50-80 (ig/L [1.9-3 (J,M]). Fish confined at sites that
were less impacted by acidification showed more moderate decrease in blood plasma Cl~
and more moderate increase in plasma Cortisol, glucose, and gill Al.
Individual Atlantic salmon physiological responses were strongly inter-correlated, with a
single principal component axis (PCI) that comprised Cortisol, glucose, Cl~, and NKA
accounting for 90% of the total variation in the physiological response variables data set
(McCormick et al.. 2009). The authors suggested that physiological impairment was the
most appropriate interpretation of PCI. In a stepwise linear regression model using water
chemistry variables only, PCI scores were best explained by low pH (r2 = 0.53). Gill Al
also was a strong predictor of physiological impairment.
Recent studies on salmon examined Al toxicity following transfer to seawater to simulate
the physiological changes associated with migration to the marine environment. In an
exposure to high acidity and inorganic Al followed by 24-hour seawater exposure, blood
plasma Cl~ levels were higher in exposed fish than in controls, suggesting reduced
seawater tolerance (Monette et al.. 2008). Loss of seawater tolerance was accompanied
by lower levels of gill NKA activity. Exposure to acidity and inorganic Al also caused
8-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
decreased plasma insulin-like growth factor (IGF-I) and levels of 3.3'.5'-triiodo-L-
thyronine (T-3). Results of this research suggest that smolt development and seawater
tolerance can be affected by exposure to high acidity and inorganic A1 despite an absence
of detectable impacts on blood plasma ion regulation in freshwater. In 2- and 5-day
Al/acid exposures of Atlantic salmon followed by a seawater challenge test, Monette et
al. (2010) showed that seawater tolerance of smolt is reduced by prior acute exposure to
acidic conditions and low levels of Al, and that the mechanisms of blood ion regulation
depend on the extent and duration of Al exposure. In both seawater and freshwater,
exposure to pH 5.3 and moderate Al levels led to accumulation of gill Al, alterations in
gill morphology, reduced gill NKA activity, and impaired ion regulation. In contrast, gill
Al accumulation was decreased, with only slight effects on gill morphology in smolts
exposed to acidic conditions and the lowest level of Al concentration.
Sockeye salmon (Oncorhynchus nerka) fry were raised in freshwater for 126 days under
sublethal conditions of low to moderate pH (4.8-6.8) by Kennedy and Picard (2012).
Effects on growth, stress, and seawater tolerance were determined after smoltification. At
the lower pH treatment (5.0), fish gained significantly less mass (average 46% of control
[pH 6.8] values), exhibited lower condition factor, and showed lower liver somatic index
values than control fish. The concentrations of liver glycogen (49% of control values) and
whole-body lipids (65% of control values) were also significantly lower. Acid exposure
caused increased stress, as measured by increased concentrations of blood plasma
Cortisol. Fish exposed to pH 5.0 in freshwater for 30 days, and then challenged with
seawater, exhibited 14% higher mortality, on average, compared to control fish. They
also showed osmoregulatory stress (increased blood plasma Na+ and CI" concentrations)
and lower critical swimming speed (22% reductions compared to control fish). Results
suggested that sockeye salmon are acid sensitive and do not acclimate to low pH under
chronic exposure conditions. This sensitivity could decrease the probability of fish
surviving after moving to the marine environment.
In a synthesis of results of bioassay studies on Atlantic salmon in Norway, gill Al
concentration was significantly correlated with water inorganic Al and ANC rFigure 8-2;
(Kroglund et al.. 2008)1. They also analyzed results of seawater challenge tests showing
that the fish had impaired hyporegulatory capacity due to inorganic Al exposure in
freshwater.
8-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
A! = aluminum; Al = inorganic monomeric aluminum; ANC = acid neutralizing capacity; dw = dry weight; g = gram; L = liter;
Lai = labile aluminum; |jg = microgram; seasurv = seawater survival.
Notes: Linear relationships are entered into the graphs whenever significant.
Source: Kroalund et al. (2008).
Figure 8-2 Relationship between (a) inorganic monomeric aluminum and gill
aluminum for parr and smolt, and (b) acid neutralizing capacity
and gill aluminum.
To assess the ability of Atlantic salmon smolts to recover from acid/Al exposure, Nil sen
et al. (2013) subjected salmon for 2- and 7-day periods to low pH (5.7) and inorganic Al
(40 ng/L [1.5 Fish were subsequently transferred to good quality water (control
exposure; pH 6.8; inorganic Al <14 jag/L [0.52 liM |). Accumulations of Al on fish gills
measured after 2 and 7 days of acid/Al exposure were 35.3 ± 14.1 and 26.6 I: 11.8 (.ig/g
(dry weight), respectively. High gill Al decreased 2 days after moving exposed fish to
control water, but gill Al was still higher than under sustained control conditions
(5-10 j_ig/L [0.18-0.37 |iM] inorganic Al) over the following 2-week period. Decreases
in blood plasma Na+ levels were observed in both test groups and remained significantly
lower than in control fish for the 2-week period after transferring fish to control water.
Blood plasma CI levels in smolts exposed for 7 days were significantly lower than in
control smolts and then remained low in both treatments following transfer to control
water. Treated smolts maintained high blood plasma glucose levels, indicative of
increased stress, after being transferred to control water and for more than a week
following exposure.
June 2018
8-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Responses of Atlantic salmon smolts to episodic pH fluctuations were assessed during
in situ experiments in rivers and streams in eastern Maine (Liebich ct al.. 201 1). Altered
plasma chloride and plasma glucose were strongly correlated with pH, consistent with the
laboratory findings of Monette and McCormick (2008) and others. High DOC
concentrations in streams could reduce the inorganic Al impact by enhancing
complexation with organic ligands.
Recently, gill Al and NKA activity were assessed in smolts migrating downstream in
New England (Kelly et al.. 2015). Smolt immigrating from the Connecticut River, where
most tributaries were well buffered, had low levels of gill Al and high levels of NKA
activity. In the Merrimack River watershed, where most tributaries were characterized by
low pH and high inorganic Al concentrations, salmon smolt immigrating downstream
showed higher gill Al and lower gill NKA activity. Stocked salmon return rates were
lower in the Merrimack River watershed, and the authors suggested that episodic
acidification could be affecting salmon smolts in poorly buffered streams in New
England.
8.3.6.2. Fish and pH Thresholds
The detrimental effects on fish associated with pH are also closely tied to ANC, which is
strongly correlated with water pH and Al [Appendix 8.3.6.4; (Driscoll et al.. 200lb')!. In
the 2008 ISA, pH effects on fish were well characterized. A pH range of 5.0 to 5.5 was
found to cause the absence of several fish species (Haines and Baker. 1986). Among
lakes with fish, there was an unambiguous relationship between the number of fish
species and lake pH, ranging from about one species per lake for lakes having pH less
than 4.5 to about six species per lake for lakes having pH higher than 6.5 (Driscoll et al..
2001b; Baker etal.. 1990b). The observations from field studies of pH effects on fish
have been corroborated by bioassay data (Figure 8-3). Some species and lifestages
experienced significant mortality in bioassays at relatively high pH (e.g., pH 6.0-6.5 for
eggs and fry of striped bass and fathead minnow; McCormick et al.. 1989; Buckler et al..
1987). whereas others were able to persist at quite low pH without adverse effects. Many
minnows and dace (Cyprinidae) are highly sensitive to acidity, but some common game
species such as brook trout, largemouth bass, and smallmouth bass are less sensitive
(threshold effects at pH <5.0 to near 5.5). In many Appalachian Mountain streams that
have been acidified by acidic deposition, brook trout is the last fish species to disappear;
it is generally lost at pH near 5.0 (MacAvov and Bulger. 1995). which usually
corresponds in these streams with ANC near 0 (ieq/L (Sullivan et al.. 2003). While brook
trout and other fish species may be absent at pH <5.0, detrimental effects on population
June 2018
8-21
DRAFT: Do Not Cite or Quote

-------
size and density may occur at higher pH values (Baker et aL 1990a: Baker and Schofield.
1985).
Critical pH Ranges of Fish
Central mudminnow
Yellow perch
Brown bullhead
Pumpkinseefi
Largemouth bass
Northern pike
Brook trou!
Whites ucker
Rock, bass
Golden shiner
Arctic char
Atlanti c sal men
Brown trout
Creek chub
Ra inbow trout
Smallmouin bass
Lake trout
Walleye
N, rebellied dace
Slimy seulpin
Common shiner
Fathead minnow
Blacknosa dace
Bluntnose minnow
Biacknose shiner
4.0
5-0
pH
7,0
Safe range, no acid-related effects occur
Uncertain range, acid relaled effects may occur
i Critical range, acid-related effects likely
Notes: Baker and Christensen (1991 generally defined bioassay thresholds as statistically significant increases in mortality or by
survival rates less than 50% of survival rates in control waters. For field surveys, values reported represent pH levels consistently
associated with population absence or loss.
Source: Fenn et al. (2011b) based on Baker and Christensen (1991).
Figure 8-3 Critical aquatic pH ranges for fish species.
June 2018
8-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Studies in the Adirondack Mountains reviewed in the 2008 ISA demonstrated the effect
of acidification on fish species richness. Of the 53 fish species recorded in Adirondack
lakes, about half (26 species) were absent from lakes with pH below 6.0. Those
26 species included important recreational species plus ecologically important minnows
that serve as forage for sport fish (Baker etal.. 1990b). There is often a positive
relationship between pH and number of fish species, at least for pH values between about
5.0 and 6.5, or ANC values between about 0 and 50 to 100 |icq/L (Cosby et al.. 2006;
Sullivan et al.. 2006a; Driscoll et al.. 2003b; Bulger etal.. 1999).
Since the 2008 review, additional pH thresholds have been published for fish species
(Table 8-2). A pH threshold of <5.9 was identified as potentially harmful to Atlantic
salmon smolts in eastern Maine (Liebich et al.. 2011). Kirbv et al. (2008) showed that
when streams in central Pennsylvania had pH below 5 because of their underlying
geology, brook trout were only present in 9 of 28 streams, whereas all streams that had
pH >6 had brook trout. Overall, the available data suggested that the threshold for brook
trout morality is at about pH = 5.0.
Despite recent reductions in acidic deposition in northern Europe, mobilized Al remains a
threat to brown trout that are native to many European fresh waters and stocked in many
U.S. waters. Andren and Rvdin (2012) identified a threshold for healthy brown trout
populations by exposing yearling trout to a pH and inorganic Al gradient in humic
streams in Scandinavia. Results suggested a threshold of less than 20 (ig/L (0.74 (iM)
inorganic Al and pH higher than 5.0. Toxic effects beyond these thresholds included Al
accumulation on the gills, increased hemoglobin and plasma glucose, decreased plasma
chloride, and increased mortality.
Brown trout embryo and first-year juvenile survival in 12 streams in northern Sweden
were investigated by Serrano et al. (2008) during snowmelt using in situ bioassays. The
study streams had high DOC, which causes a pH decrease, but also protects fish against
Al toxicity. High juvenile brown trout mortality was documented during the spring flood,
in association with low pH. No significant effect of inorganic Al concentration on
juvenile or embryo mortality was observed, likely due to Al complexation by organic
acid anions. An empirical model developed to predict juvenile brown trout mortality in
high-DOC streams suggested a critical chemical threshold of pH in the range of 4.8-5.4.
High embryo and yolk sac fry survival was recorded, even at stream sites with pH as low
as 4.0. As expected, the observed pH threshold in DOC-rich water was lower than
previously observed thresholds for low-DOC freshwaters. First-year juveniles are likely
to be most sensitive to adverse effects of low pH in northern boreal stream ecosystems,
especially during snowmelt-driven episodic pH depressions.
8-23
DRAFT: Do Not Cite or Quote

-------
Table 8-2 pH thresholds in fish published since the 2008 Integrated Science
Assessment for Oxides of Nitrogen and Sulfur-Ecological Criteria.
Fish Species and Effects
Region
Potential Thresholds
Reference
Atlantic salmon smolts reduction of plasma
ions
Eastern Maine
pH 5.9
Liebich et al. (2011)
Brook trout loss of whole-body Na
Great Smoky
Mountains NP
pH 5.1
Neffet al. (2008)
Brook trout loss of whole-body Na of 10 to
20%
Great Smoky
Mountains NP
pH 4.9 to 5.1
Neffet al. (2009)
Juvenile brown trout mortality in high DOC
streams
Sweden
pH 4.8 to 5.4
Serrano et al. (2008)
Brown trout embryo and yolk sac fry
survival during episodes in DOC-rich lakes
Sweden
pH 4.0
Serrano et al. (2008)
Toxicity to brown trout in humic streams
Northern Europe
pH 5.0; inorganic
aluminum 20 |jg/L
Andren and Rvdin (2012)
Response of Atlantic salmon to alarm cues
after episodic exposure3
Canada
6.2
Leduc etal. (2009)
Interference of chemical alarm cues to
assess predation risk in juvenile Atlantic
salmon3
Canada
pH <6.6
Elvidae and Brown
(2014)
DOC = dissolved organic carbon; Na = sodium; NP = national park.
aBehavioral responses offish are discussed in Appendix 8.3.6.5. Table B-23 and Table B-24 in the 2008 ISA summarize pH
thresholds from the NAPAP report (Baker et al.. 19903).
8.3.6.3. Fish and Acid Neutralizing Capacity Thresholds
1	ANC has been found in various studies reviewed in the 2008 ISA to be a good single
2	indicator of the biological response and health of aquatic communities in acid-sensitive
3	systems (U.S. EPA. 2008a; Sullivan et al.. 2006a). For fish and other aquatic biota, ANC
4	is closely tied to pH (Appendix 8.3.6.2) and the bioavailability of Al [Appendix 8.3.6.4;
5	(Driscoll et al.. 200lb) 1. There is often a positive relationship between pH or ANC and
6	number of fish species, at least for pH values between about 5.0 and 6.5, or ANC values
7	between about 0 and 50 to 100 (j,eq/L (Cosby et al.. 2006; Sullivan et al.. 2006a; Bulger et
8	al.. 1999). In Shenandoah National Park streams in Virginia, fish species richness was
9	lower by one species, on average, for every 21 (j,eq/L decrease in ANC (Sullivan et al..
10	2003; Bulger etal.. 1999). Interpretation of species richness can be difficult, however,
June 2018
8-24
DRAFT: Do Not Cite or Quote

-------
1	because more species tend to occur in larger lakes and streams as compared with smaller
2	ones, irrespective of acidity (Sullivan et al.. 2003). This might be due to increased aquatic
3	habitat complexity and diversity in larger watersheds (Sullivan et al.. 2003).
4	As summarized in the 2008 ISA, lakes and streams having an annual average
5	ANC < 0 (.ieq/L generally do not support fish (Figure 8-4). The analysis shown in this
6	figure suggests that there could be a loss of fish species with decreases in ANC below a
7	threshold of approximately 50 to 100 ficq/L (Sullivan et al.. 2006a).
cfl
0)
"o
0)
a
W
.e
cfl
LL
*¦
O
i_
a
-200
-100
100 200
ANC (/jeq/L)
300
400
500
ANC = acid neutralizing capacity; L = liter; peq = microequivalent.
Notes: The data are presented as the mean (filled circles) of species richness within 10 jjeq/L ANC categories, based on data
collected by the Adirondacks Lakes Survey Corporation.
Source: Modified from Sullivan et al. (2006a).
Figure 8-4 Number of fish species per lake verses acidity status, expressed
as acid neutralizing capacity, for Adirondack lakes.
Various ecological effect criteria, including effects on fish communities, have been used
to classify ranges of ANC values (Table 8-3). The use of these criteria lies in the
June 2018
8-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
association between ANC and the surface water constituents that directly affect
acidity-related stress. These include, in particular, pH, Ca2+, and inorganic A1
concentration. Bulger et al. (2000) developed ANC thresholds for brook trout response to
acidification in forested headwater catchments in western Virginia. Across the eastern
U.S., brook trout are often selected as a biological indicator of aquatic acidification
because they are native to many eastern surface waters and because residents place
substantial recreational and aesthetic value on this species. Note that because brook trout
are comparatively acid tolerant, adverse effects on many other fish species would be
expected at higher ANC than the threshold values that impact brook trout. Annual
average ANC greater than about 50 j^icq/L is generally considered suitable for brook trout
in southeastern U.S. streams. Such streams have sufficient buffering capacity to prevent
acidification from eliminating this species, and there is reduced likelihood of lethal
storm-induced acidic episodes. In streams having ANC > 50 (ieq/L, reproducing brook
trout populations are expected if the habitat is otherwise suitable (Bulger et al.. 2000). but
some streams may periodically experience episodic chemistry that affects species more
sensitive than brook trout. Streams having annual average ANC from 20 to 50 j^ieq/L may
experience episodic acidification during storms to pH and ANC levels that can be lethal
to brook trout, as well as other fish (Bulger et al.. 2000). Streams that are designated as
episodically acidic (chronic ANC from 0 to 20 j^ieq/L) are considered marginal for brook
trout because acidic episodes are likely (Hveretal.. 1995). although the frequency and
magnitude of episodes can vary widely. Streams that are chronically acidic (average
ANC less than 0 j^icq/L) are not expected to support healthy brook trout populations
(Bulger et al.. 2000).
Hesthagen et al. (2008) used a regional water chemistry database to determine critical
values of pH, ANC, and inorganic Al for survival of brown trout in 790 Norwegian lakes.
The threshold value (ANCimut) to avoid fish damage was compared with that found in a
similar study conducted in 1986. In 1995, the threshold ANC value to avoid toxic effects
to fish and retain unaffected fish populations was 67 (ieq/L, compared with 20 j^ieq/L in
1986. The higher ANCimut found for 1995 was attributed to lower pH and higher
inorganic Al concentration at a given ANC value in 1995 than in 1986. This, in turn, was
attributed to increases in total organic carbon in these lakes. Thus, the value of ANC as
an indicator of biological effects may differ between acidification and recovery periods.
8-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
Table 8-3 Expected ecological effects and concern levels in freshwater
ecosystems at various levels of acid neutralizing capacity.
Category Label
ANC Level
|jeq/L
Expected Ecological Effects
Low concern
(no effect)
>100
Fish species richness may be unaffected. Reproducing brook trout populations
are expected where habitat is suitable. Zooplankton communities are unaffected
and exhibit expected diversity and distribution.
Moderate concern
(minimally
impacted)
50 to 100
Fish species richness begins to decline (sensitive species are lost from lakes).
Brook trout populations are sensitive and variable, with possible sublethal
effects. Diversity and distribution of zooplankton communities begin to decline as
species that are sensitive to acid deposition are affected.
Elevated concern
(episodically acidic)
Oto 50
Fish species richness is greatly reduced (more than half of expected species are
missing). On average, brook trout populations experience sublethal effects,
including loss of health and reproduction (fitness). During episodes of high
acidity, brook trout may die. Diversity and distribution of zooplankton
communities decline.
Acute concern
<0
Near complete loss offish populations is expected. Planktonic communities have
extremely low diversity and are dominated by acid-tolerant forms. The numbers
of individuals of plankton species that are present are greatly reduced.
ANC = acid neutralizing capacity; L = liter; |jeq = microequivalent.
Based on data from southern Appalachian steams and from Shenandoah National Park.
Source: based on U.S. EPA (20090).
8.3.6.4. Fish and Aluminum Thresholds
The detrimental effects of A1 on fish are closely tied to A1 solubility, complexation, and
speciation, which are all strongly influenced by pH (Appendix 8.3.6.2). ANC
(Appendix 8.3.6.3). and concentrations of organic acids (Driscoll et al.. 2001b). A1 has no
established biological function and dissolved inorganic Al can be highly toxic to fish and
other aquatic biota. The current U.S. EPA acute and chronic Aquatic Life Criteria for Al
are 750 and 87 (ig/L, respectably, and not more than once every 3 years on average (U.S.
EPA. 1988). In the 2008 ISA, elevated concentrations of inorganic Al associated with
acidification of surface waters were shown to affect fish populations and communities in
parts of the Adirondack Mountains of northern New York (Simonin et al.. 1993; Kretser
and Gallagher. 1989; Johnson et al.. 1987; Schofieldand Driscoll. 1987; Baker and
Schofield. 1982). in acid-sensitive streams of the Catskill Mountains of southeastern New
York (Charles and Christie. 1991). and the Appalachian Mountains from Pennsylvania to
Tennessee and South Carolina (Bulger et al.. 2000; Bulger etal.. 1999; SAMAB. 1996).
In one study reviewed in the 2008 ISA, 20% mortality of caged young-of-year brook
June 2018
8-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
trout in poorly buffered headwater streams in the Adirondacks was documented during a
30-day period with a median inorganic A1 concentration of 54 (ig/L (2 |iM/L: Baldigo ct
al.. 2007). The authors estimated that 90% mortality would occur over 30 days, in
response to a median inorganic Al concentration of 108 (ig/L (4.0 (iM/L). Threshold
values for Al from Baker et al. (1990a) for various species and effects were summarized
in the 2008 ISA. These values, along with additional Al thresholds for various fish
species and endpoints, are listed in Table 8-4. Field studies of the effects of Al on aquatic
biota are typically confounded by simultaneous pH effects. Both inorganic Al and H+ can
be toxic, and the solubility of inorganic Al is strongly pH-dependent.
Since the 2008 ISA, additional Al thresholds for salmonids have been published
(Table 8-4). In a synthesis of 347 short term (<14 days) exposures (in tanks fed river
water or in tanks where water quality was manipulated) of salmon parr and smolt
performed between 1990 and 2003 in Norway, Kroglund et al. (2008) identified
dose-response relationships for pH, inorganic Al concentration, ANC, gill Al, and time of
first fish mortality over a 10-day exposure period (Figure 8-5). Results of smolt releases
were also evaluated after pre-exposure to moderately acidic waters. All smolt survived at
pH >5.8 and inorganic Al <200 (ig/L (7.4 (iM). For parr, mortality increased at pH <5.6
or inorganic Al >45 (ig/L (1.7 (iM). In the same study, seawater challenge tests were
analyzed where smolts were pre-exposed to sublethal concentrations of Al in freshwater
and then moved to saltwater. As freshwater pH decreased and inorganic Al concentration
increased, hypo-osmoregulatory capacity decreased, and increased mortalities were
observed. Smolt exposed to >5 to <10 (ig/L (0.18-0.37 |iM) inorganic Al in freshwater
and then released into seawater had a 25 to 50% reduction in survival based on data from
a sea survival program in the River Imsa in Norway (Kroglund et al.. 2008; Kroglund et
al.. 2007).
Additional studies have characterized the role of DOC in influencing Al bioavailability
and subsequent effects on fish. In a field study of Atlantic salmon smolts in streams in
eastern Maine, observations suggested that streams with moderate to high DOC
concentrations buffered the effects associated with low pH and high total Al
concentrations (Liebich et al.. 2011). Smolts exposed to high DOC waters had lower gill
Al concentrations compared to smolts in low DOC waters having similar pH and Al
concentrations.
8-28
DRAFT: Do Not Cite or Quote

-------
Table 8-4 Threshold values of aluminum for various fish species and associated effects.
Type of Study
Taxa
PH
Al (Mg/L)
Observed Effect
Form of Al
Country
Reference
Field study
Brook trout
(Salvelinus
fori tin alis)
4.9
286
No survival of trout stocked into
lakes with higher total Al (even
after accounting for pH effects)
Total
U.S.
Schofield and Troinar
(1980)
Laboratory
exposure
White sucker
(Catostomus
commersoni)
5.2
200
>50% larval mortality
Total
U.S.
Baker and Schofield (1982)
Laboratory
exposure
Brown trout
(Salmo trutta)
4.5-5.4
250
>50% fry mortality
Total
-
Brown (1983)
Laboratory
exposure
Atlantic salmon
(Salmo salar)
4.9-5.0
130
Significant increase in mortality
of presmolts
Labile
Norway
Rosseland and Skoaheim
(1984)
Laboratory
exposure
Eel (Anguilla
anguilla)
5.1
230
Significant increase in elver
mortality
Total
Norway
Fiellheim et al. (1985)
Field mesocosm
experiment
Atlantic salmon
(Salmo salar)
5.1
75
>50% mortality of smolts
Labile
Norway
Skoaheim and Rosseland
(1986)
Field survey
Brown trout
(Salmo trutta)
-
40
Fish absent or rare in streams in
Wales and England
Labile
monomeric
Wales and England
Turnpenny et al. (1987)
Laboratory
exposure
Blueback
herring (Alosa
aestivalis)
5.5-5.6
100
>50% larval mortality
Total
U.S.
Klauda and Palmer (1987)
Whole-stream
experiment
Atlantic salmon
(Salmo salar)
and brown trout
(Salmo trutta)
5.0
350
>50% mortality of young-of-the-yr

Wales
Ormerod et al. (1987)
Laboratory
exposure
Brown trout
(Salmo trutta)
5.2
30
Significant reduction in fish
growth
-
England
Sadler and Lvnam (1988)
June 2018
8-29
DRAFT: Do Not Cite or Quote

-------
Table 8-4 (Continued): Threshold values of aluminum for various fish species and associated effects.
Type of Study
Taxa
PH
Al (Mg/L)
Observed Effect
Form of Al
Country
Reference
Laboratory
exposure
Walleye
(Stizostedion
vitreum)
4.9
50
>50% mortality of embryos to
4-days post-hatch
Total
Canada
Holtze and Hutchinson
(1989)
Laboratory
exposure
Brook trout
(Salvelinus
fori tin alis)
5.2
29
Survival of 1-yr-olds decreased
Inorganic
monomeric
U.S.
Inqersoll et al. (1990)
Laboratory
exposure
Brook trout
(Salvelinus
fon tin alis)
4.8
34
Weight of 1-yr-olds decreased
after exposure to either pH or Al
threshold
Inorganic
monomeric
U.S.
Inqersoll et al. (1990)
Laboratory
exposure
Greenback
cutthroat trout
(Oncorhynchus
clarki stomias)
5.5
100
Decreased survival of alevins
and swim-up larvae
Total
U.S.
Woodward etal. (1991)
Laboratory
exposure
Greenback
cutthroat trout
(Oncorhynchus
clarki stomias)
5.5
100
Decreased swimming activity in
swim-up larvae and alevins
Total
U.S.
Woodward etal. (1991)
Laboratory
exposure
Greenback
cutthroat trout
(Oncorhynchus
clarki stomias)
5.0
300
Decreased survival of embryos
Total
U.S.
Woodward etal. (1991)
Laboratory
exposure
Greenback
cutthroat trout
(Oncorhynchus
clarki stomias)
5.0
300
Increased loss of Na ions in
embryos
Total
U.S.
Woodward etal. (1991)
Laboratory
exposure
Greenback
cutthroat trout
(Oncorhynchus
clarki stomias)
6.0
50
Increased loss of Na, K, and Ca
ions in alevins (compared to
treatment with same pH and
without Al)
Total
U.S.
Woodward etal. (1991)
June 2018
8-30
DRAFT: Do Not Cite or Quote

-------
Table 8-4 (Continued): Threshold values of aluminum for various fish species and associated effects.
Type of Study
Taxa
PH
Al (Mg/L)
Observed Effect
Form of Al
Country
Reference
Laboratory
exposure
Greenback
cutthroat trout
(Oncorhynchus
clarki stomias)
6.0
50
Decreased feeding strikes in
swim-up larvae (compared to
treatment with same pH and
without Al)
Total
U.S.
Woodward etal. (1991)
Laboratory
exposure
Greenback
cutthroat trout
(Oncorhynchus
clarki stomias)
6.0
50
Decreased swimming duration in
alevins (compared to treatment
with same pH and without Al)
Total
U.S.
Woodward etal. (1991)
Field study
Blacknose dace
(Rhinichthys
atratulus)
-
101
Decreased survival of adults
when exposed to threshold for
more than 6 days
Inorganic
monomeric
U.S.
Simonin et al. (1993)
Field study
Brook trout
(Salvelinus
fon tin alis)
-
225
Juvenile mortality significantly
increased (>20%) when exposed
to Al threshold for 2 or more days
Inorganic
monomeric
U.S.
Baldiao and Murdoch
(1997)
Field study
Brook trout
(Salvelinus
fon tin alis)

54/108
Correlations between low
(54 ijg/L) and high (108 pg/L)
thresholds and low (20%) and
high (50-90%) mortality
Inorganic
monomeric
U.S.
Baldiao and Murdoch
(1997)
Laboratory
exposure
Atlantic salmon
(Salmo salar)
5.0-5.4
43-68
Ion regulation, stress physiology,
gill Al accumulation
Inorganic
U.S.
Monette and McCormick
(2008)
Laboratory
exposure
Atlantic salmon
(Salmo salar)
5.4-6.3
28-64
Ion regulation, seawater
tolerance, plasma hormone
levels, stress physiology
Inorganic
U.S.
Monette et al. (2008)
Laboratory
exposure
Atlantic salmon
(Salmo salar)
5.8
40
Smolt survival
Cationic
(inorganic
aluminum)
Norway
Kroalund et al. (2008)
Laboratory
exposure
Atlantic salmon
(Salmo salar)
5.6
45
Parr survival
Cationic
(inorganic
aluminum)
Norway
Kroalund et al. (2008)
June 2018
8-31
DRAFT: Do Not Cite or Quote

-------
Table 8-4 (Continued): Threshold values of aluminum for various fish species and associated effects.
Type of Study
Taxa
PH
Al (Mg/L)
Observed Effect
Form of Al
Country
Reference
Field bioassay
Atlantic salmon
(Salmo salar)
5.4-5.6
50-80
Survival, smolt development, ion
homeostasis, stress
Inorganic
U.S.
McCormick et al. (2009)
In situ exposure
Brown trout
(Salmon trutta)
5.0
20
Recommended threshold to
sustain healthy populations
based on gill Al, blood physiology
Inorganic
Sweden
Andren and Rvdin (2012)
Laboratory
exposure
Atlantic salmon
(Salmo salar)
5.7
40
Gill Al accumulation
Inorganic
U.S.
Nilsen et al. (2013)
June 2018
8-32
DRAFT: Do Not Cite or Quote

-------
o LAI; smolt * Ali; smelt
*
i
y
o
<
1D0
90
80
70
60
50
40
30
20
10
0

r-**-
few
mmrn
4.5
5.5
6.5
pH
o LAI; smolt » Ali; smalt
i
<15
¦e
o
Ł
e
=
o
o
<
100
90
30
70
60
50
40
30
20
10
0
-i O ¦ 0- » fln


0


D


O


&


*


* o


m



P

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
8.3.6.5.
Behavioral Responses to Acidification
New studies in salmonids have shown effects of acidification on behavioral endpoints.
Using hatchery-reared juvenile Atlantic salmon, Elvidge and Brown (2014) conducted a
tethering experiment in reaches of neutral (pH >6.6) and low pH (<6.6) salmon nursery
streams, plus one additional stream that varied between pH classes. The researchers
observed that the tethered fish in the low-pH streams were more likely to be preyed upon
than fish in the neutral streams although there were fewer predatory fish species, similar
availability of physical refugia, and similar threat from terrestrial predators. Leduc et al.
(2009) quantified alarm behavior of juvenile Atlantic salmon exposed to chemical alarm
cues in a stream before and after rainfall. Before rainfall, fish exhibited an alarm response
in the study streams. After rainfall, fish from the higher pH (near 7.5) stream continued to
respond, but those from the lower pH (near 6.2) stream did not. The researchers
suggested that episodic acidification in small nursery streams may disrupt the information
mediated by chemical alarm cues and contribute to higher fish mortality. This may be the
first published study documenting the disruption of a chemical messenger by
acidification of stream water.
8.3.7. Fish-Eating Birds
The 2008 ISA reviewed the limited studies that documented effects of acidification on
birds. Acidification has been shown to disrupt food web dynamics causing alteration to
bird diet, breeding distribution, and reproduction. Lacoul et al. (2011) reviewed studies of
acidification on aquatic birds and noted that in the Atlantic Canada region, breeding was
limited to lakes and streams that had pH higher than 5.5. A substantial amount of
research has recently been conducted on air pollution effects on birds, but the focus of
that work has been on mercury exposure, and results are not discussed here. Deposition
effects of mercury on biota in the Adirondacks are described in Appendix 12.
8.3.8. Aquatic Assemblages
In the 2008 ISA, decreases in species richness in response to surface water acidification
were reported in lakes and streams for all major trophic levels of aquatic organisms
(Baker etal.. 1990a). even after adjusting for lake size (Matuszekand Bcggs. 1988;
Schofield andDriscoll. 1987; Frenette et al.. 1986; Rago and Wiener. 1986; Harvey and
Lee. 1982). Acidification was also found to cause decreases in food web complexity
(indicated by the number of trophic links or species) in Adirondack lakes (Havens and
Carlson. 1998).
June 2018
8-34
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Since the 2008 ISA, several studies have considered effects of species assemblages
associated with acid-impacted water bodies. Nierzwicki-Bauer et al. (2010) presented
baseline midsummer chemistry and community composition of phytoplankton, bacteria,
rotifers, crustaceans, macrophytes, and fish in 30 lakes of the southwestern Adirondacks
(Figure 8-6). Species richness in the lakes was correlated with acid-base chemistry at all
trophic levels except bacteria. The decrease in number of taxa per unit pH was similar
among groups and ranged from 1.75 (crustaceans) to 3.96 (macrophytes).
Following the 5-year experimental acidification of Little Rock Lake in Wisconsin from
pH 6.1 to 4.7, Hogsden et al. (2009) observed greater reductions in species richness at
higher trophic levels (fish and zooplankton) compared to algae. The authors suggested
that the observed asymmetrical food web response could have implications for biological
recovery of lakes. This study was limited, however, by the fact that the experimental
acidification only involved the lake and not the watershed. Cladoceran assemblages
reconstructed by Mosscrop et al. (2015) from lake sediment cores from 30 lakes in
Ontario, Canada showed that in shallow lakes, there was no significant difference in the
remains of cladoceran taxa deposited in sediments between present-day and preindustrial
sediments. In deeper lakes, relative abundance of cladocerans had shifted overtime. The
researchers hypothesized that cladoceran assemblages in shallower lakes might be
influenced more by habitat than by water chemistry and that Ca availability may vary
spatially in shallow lake areas or that cladoceran species living in littoral zones may be
less sensitive to low Ca.
In an analysis of pH effects on grazing and herbivory in 20 stream food webs across the
British Isles, considerable species differences were observed across a pH gradient from
5.0 to 8.4 (Layer et al.. 2013). Three dose-response relationships were characterized for
diatom assemblages (Figure 8-7). Macroinvertebrate taxon richness and benthic density
were plotted against stream pH for all primary consumer taxa as well as by trophic group
(shredders, herbivore-detritivores, collectors, grazers; Figure 8-8). At low pH, generalist
herbivore-detritivores dominated the primary consumer assemblage. At higher pH, they
were partially replaced by specialist grazers. The researchers concluded that the ability of
acid-tolerant herbivore-detritivores to exploit food resources and the low nutritional value
of basal resources in acidified streams might decrease recolonization by specialist grazers
in streams where chemical conditions are becoming less acidic.
8-35
DRAFT: Do Not Cite or Quote

-------
Bactcrioplankton
~ ~
y=2.4972x - 0.1353
~ ~ ~
R'=0.049
~
I,

	
— «i ¦ " ¦ ¦
w m ¦
¦
y=1.4259x-1.9141 —
R'-0.122
•125 4 75 525 5 75 6.25 6,75 7 25 7 75
Rotifers
y^3,S603x - 12.249
R*-0.6863
•<25 4 75 5.25 5.75 6,25 675 7.25
~ ~
Macrophytes
~
y-3.957x • 9.3734
~ ~
R-—0.2128
~ ~ ~
—— i ¦ *—

TV.—- ~ ~
~ ~
	 ^

~ ~*
K


Phytoplanklon
y=3.9744x - 9.0623
~ ~
*R'-0.49%
~
~ .


~ ~ *5^-1
~
~

42)	4 75	5 25	575	625	6,75	7 25
Crustaceans
30
25
20
15 |
10
5 1-
y-l,7529x- 3.4791
R-0.5873
4,25 4.75 5.25 5.75 6.25 6.75 7.25
Fish

y=3.7218x - 17-641
R!=0.5316
~
~
~
A r~
	—
425 4.75 5.25 5,75 625 6.75 725
Lake pH
Notes: The y-axis denotes number of groups (bacteria), taxa (phytoplankton) or species (rotifers, crustaceans, macrophytes, fish).
The x-axis denotes the pH range that occurred in the 30 study lakes. Note; in the bacterioplankton graph, two regressions were run
on the basis of genera (diamonds) and classes (squares).
Source: Nierzwicki-Bauer et ai. (2010't.
Figure 8-6 Species richness of biotic groups in 30 Adirondack study lakes
relative to midsummer epilimnetic pH during sample years.
June 2018
8-36
DRAFT: Do Not Cite or Quote

-------
100 1
1
CL
a
10 -
z
4:
u
©
CL
CO
0
O
O
o
y=5.37X-20.3;
r2=0.47; F=15.84;
p=0.001*
10
s
a
—
10 10 -
106.
o
o
y=2.2*1012X-1.26*1013;
r5=0.45; F=14.49;
p=0.001*
XT
CD
E

-------
3 All primary consumer taxa
4 i
, Q-
o

%
y=0.17»O.D4, r-030, F=7JZ
p=0.012"
C Shretiders
y=0.24*-0.9B, r^=0.39, F=11.30.
p=0.«H"
1 -
0 -I—OC—O—r
e He?bivore-de!ri1ivores
>t-O.CSx+(5.B2. r'=0.09. F=1 86.
p=0.19 (n.B.)
<3 ^ 8
o oo> o o
O OX)
g Collectors
>=0.t 1*+tl.10, rr=0.14; F-2.B4;
p=a.n (n.s.)
V
o o
o
I Grazers
)fc0.26j(-1 ,2a. 1^=0.43. F=13.37,
p-0.002'
b All primary consumer 1axa
O
y=037x*Q2, P^).24, F=5.70.
p=Q.O20"
d Shredders
y=0 94jsl 65, r"=0.67, F=36. j
ptO.0001*	O
E 4
'to
6
o o ~
f Herbwore-decrilwores
y=0-08x+1.15, P=Q.G1. F=0.25,
p=0.63 (n.s.)
o
o°°
h Collectors
O ° o ° O
00° ° o
U Q)
y=0.31 *1-021, ^=0.15; F=3.11;
p=O.Q95 (n.S.)	
j Grazers
y=0.76*-3-9. ^=0.53. F=20.00,
pcD.001 *
O
O
°o
pH
ri.s. = not significant; S = sulfur.
Notes: statistical significance was determined using linear regression analysis. Asterisk denotes a statistically significant result at
p < 0.05.
Source: Layer et al. (2013).
Figure 8-8 Macroinvertebrate community composition in 20 streams across
a pH gradient. Taxon richness (total number of primary consumer
taxa; left panel) and benthic density [number of individuals
logio(x+ 1)-transformed; right panel] plotted against stream pH,
all primary consumers (a, b), shredders (excluding Leuctridae and
Nemouridae) (c, d), herbivore-detritivores (Leuctridae and
Nemouridae) (e, f), collectors (g, h) and grazers (i, j).
June 2018
8-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Lacoul etal. (2011) reviewed available information on the effects of water acidification
on aquatic organisms in Atlantic Canada. The pH threshold 5.0 was critical to sensitive
macrophytes.
New analytical methods have recently been developed for detecting and quantifying
ecological thresholds, such as tipping points for ANC or other acid-base chemical
parameters at which adverse impacts on aquatic assemblages occur (Baker and King.
2010). Previous methods for designating ecological community thresholds have been
based on univariate indicators or multivariate dimension-reduction of community
structure. They tend to be insensitive to the responses of taxa that have low occurrence
frequencies or highly variable abundance. The Threshold Indicator Taxa Analysis
(TITAN) statistical tool detects changes in taxa distributions and abundances along an
environmental gradient, such as ANC for example. It uses indicator species scores to
integrate occurrence, abundance, and directionality of taxa responses by optimizing the
value of a continuous variable that partitions sample units while maximizing
taxon-specific scores. It emphasizes the relative magnitude of change and increases the
contributions of taxa that have low occurrence frequencies and high sensitivity to the
gradient. As an example, Baker and King (2010) assessed macroinvertebrate community
response to a nutrient (e.g., P) gradient using TITAN. Results supported previous
threshold estimates but with lower confidence limits. Several taxa were shown to decline
at lower levels of P concentration.
In 28 subwatersheds of the Neversink River Basin in the Catskill Mountains, NY,
Harpold et al. (2013) used a simple hydrogeomorphic model as a tool to predict impacts
of acidification on biological communities. The model, which considered watershed slope
and drainage density to estimate ANC, successfully predicted several measures of
macroinvertebrate and fish community richness.
8.4. Documentation of Biological Recovery
The biological impact of acidifying deposition is mediated through changes in water
quality that in turn impact biota. Deposition of N and S can lead to a cascade of
biogeochemical changes (Appendix 4 and Appendix 7) that may induce biological
effects. It is often difficult to separate nutrient enrichment (Appendix 9) and acidification
stressors, as they act simultaneously in some water bodies. Poschetal. (2011) argued that
linking nutrient and acidity chemical criteria to vegetation occurrence on a regional scale
is an approach that can be used to identify optimal N and S deposition limits to sustain a
prescribed biodiversity goal (see Appendix 8.5.4).
June 2018
8-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Trajectories of biological recovery as used in this ISA are defined in Chapter 1.11.1.
Complete biological recovery would entail a return to similar species make-up, richness,
and abundance as existed in the ecosystem in question prior to the advent of
human-caused acidic deposition (around the year 1860). Full recovery might be defined
as restoring healthy populations consistent with regional expectations. Biological
recovery from past acidification is a process affected by chemical, climatic, biological,
chance, and hydrologic influences over time. It may, under certain conditions, follow
chemical recovery of such water quality constituents as pH, ANC, and the concentrations
of SO42 , NO3 . inorganic Al, and DOC. Both chemical and biological recovery can, and
often do, lag behind changes in the levels of S and/or N emissions and deposition because
of chemical, hydrological, and biological processes and constraints. In addition, the ANC
level that reflects recovery of pH or inorganic Al may differ between the acidification and
recovery phases (Hcsthagcn et al.. 2008). In a practical sense and depending on the level
of impact, complete biological recovery may not be attainable at many acidified locations
within a reasonable management timeframe because soil reserves of base cations at many
locations have been depleted in response to many decades of acidic deposition and
because other stressors, in addition to acidic deposition, have also altered ecosystem
structure and/or function or will do so in the coming decades. Such stressors include
changes in climate, land use, and other perturbations. More commonly, only partial
biological recovery may be possible. Ecosystems deemed to be on a recovery trajectory
are those found to be moving towards a mix of species presence and abundance that
approximates the undisturbed state.
Biological recovery can occur only if chemical recovery (Appendix 7) is sufficient to
allow growth, survival, and reproduction of acid-sensitive plants and animals (Driscoll et
al.. 2001b). Also, chemical recovery of ANC or pH may not necessarily follow the
reverse of the acidification path due to changes in relationships among ANC, pH, DOC,
and Al; depletion of soil base cation pools; hydrology; climate; and/or partially reversible
(or irreversible) S adsorption on soils. In the 2008 ISA, studies of biological recovery
generally indicated that the time required for biological recovery is uncertain and that
biological responses typically lag behind chemical recovery and may take decades from
the onset of chemical recovery (U.S. EPA. 2008a).
Literature reviewed in the 2008 ISA suggested that the time required for recovery of biota
varies. In general, macroinvertebrate populations in streams recover more rapidly (within
approximately 3 years in response to improved chemical conditions), relative to lake
populations of zooplankton, followed by increases in lake fish populations which may
occur 5 to 10 years after zooplankton recovery (Driscoll et al.. 2001b; Gunn and Mills.
1998). Biological recovery of previously acidified surface waters can lag behind chemical
recovery because of factors such as limits on dispersal and recolonization (Gray and
8-40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Arnott. 201 1). barriers imposed by water drainage patterns (Jackson and Harvey. 1995).
the influence of predation (Layer etal.. 2011; McNicol et al.. 1995). and other
environmental stressors (Yan et al.. 1996b; Yan et al.. 1996a; Gunn et al.. 1995; Havas et
al.. 1995; Jackson and Harvey. 1995; McNicol et al.. 1995). Biological recovery is likely
to occur in stages due to differences among species in acid sensitivity and the rate of
recovery of affected organisms (Driscoll et al.. 2007b'). In a study reviewed in the 2008
ISA, recovery of the zooplankton community in an experimentally acidified lake did not
retrace the same trajectory as the initial deterioration due to acidification, indicating
substantial hysteresis in recovery (Frost et al.. 2006). Biological recovery research from
the Sudbury region of Ontario, Canada was highlighted in the 2008 ISA. This region,
formerly impacted by smelter operations, has been important for clearly documenting the
chemical and biological effects of S and metal deposition as well as subsequent recovery
from acidification.
New studies at multiple trophic levels continue to support findings presented in the 2008
ISA that biological recovery has lagged behind chemical recovery in many systems and
that the lag response may vary among taxa and water bodies. Several long-term studies of
water acidification indicate that biological recovery has been limited despite significant
deposition reductions and improvements in water chemistry (Baldigo et al.. 2016;
Battarbee et al.. 2014; Murphy et al.. 2014; Angeler and Johnson. 2012). Other studies,
such as Honnedaga Lake and Brooktrout Lake in the Adirondacks, show more evidence
for return of biota to levels approaching preacidification levels (Sutherland et al.. 2015;
Josephson et al.. 2014). In Brooktrout Lake, biological recovery of the food web structure
has begun, in part, due to reintroduction and reestablishment of brook trout in the lake.
Ongoing biological recovery cannot necessarily be expected to conclude with the return
of the biological community to preacidification conditions, however. This is due to
factors such as differences in the rate of recovery of aquatic organisms, permanent loss of
some acid-sensitive species, and irreversible chemical and physical alterations to aquatic
environments during acidification (NAPAP. 2011). Effects of other stressors can modify
the recovery trajectories of aquatic ecosystems that have experienced acidification. Of
particular importance in this regard are the introduction of species of fish or other
organisms, either purposely as in the case of fish stocking by state agencies or local
organizations or inadvertently. In the coming years, warming of surface waters in
response to climate change may become increasingly important. For example,
(McDonnell et al.. 2015) showed that stream acidity (a top-down stress) together with
stream warming (a bottom-up stress) cause a reduction in the extent of stream habitat in
the southern Appalachian Mountains that is suitable for cold-water aquatic biota.
8-41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
8.4.1.
Phytoplankton Recovery
In the 2008 ISA, phytoplankton recovery from experimental acidification showed an
increase in phytoplankton species richness and diversity with increased lake pH. In
Lake 223 in the Experimental Lakes area of Ontario, there was little increase in
phytoplankton diversity as pH changed from 5.0 to 5.8 but a strong recovery of diversity
at pH above 6 (Findlav and Kasian. 1996). In Lake 302S, profound biological changes
began at pH 5.5, with phytoplankton assemblages at pH below 5.5 resembling those in
acidified lakes. Findlav (2003) reported that lakes that were previously low in pH (5.0 to
5.5) and are now above pH 6 have shifted towards phytoplankton assemblages more
typical of circumneutral environments. Recent studies on responses of phytoplankton to
decreased acidity have been limited. Josephson et al. (2014) reported that observed
increases in chlorophyll a and decreases in water clarity in Honnedaga Lake, NY in
recent years reflect an increase in phytoplankton abundance in association with chemical
recovery from acidification.
The 2008 ISA reported results of several paleolimnological studies that used fossil
remains of diatoms or chrysophytes in lake sediments to infer lake chemistry at discrete
time periods in the past. These studies provided valuable information regarding the extent
of historical acidification, especially in the Adirondack Mountains, by using
well-established relationships between water chemistry and diatom community structure.
The studies also provided the foundation for testing the performance of process-based
dynamic models, such as Model of Acidification of Groundwater in Catchments
(MAGIC), by comparing MAGIC hindcast simulations of water chemistry with diatom
inferences of preindustrial pH and ANC (Sullivan et al.. 1990).
Since the 2008 ISA, a number of additional paleolimnological studies have been
conducted that have documented historical acidification and subsequent recovery
(Table 8-5); however, most of the studies have been conducted in other countries. In one
study from the U.S., diatom shifts were linked to historical changes in pH at Brooktrout
Lake in the Adirondacks. Fragilariforma acidobiontica, a diatom that is often abundant
at pH <5.0, was present in lake sediments deposited since the 1950s and shifts in
Mallomonas and Synura were also observed (Sutherland et al.. 2015). In another study,
Arseneau et al. (2016) concluded that Adirondack lakes that were not previously acidified
by acidic deposition will likely not recover to predisturbance chrysophyte community
structure because of the influence of other stressors, including changes in climate.
June 2018
8-42
DRAFT: Do Not Cite or Quote

-------
Table 8-5 Paleolimnological responses to changing lake chemistry published
since the 2008 Integrated Science Assessment for Oxides of
Nitrogen and Sulfur-Ecological Criteria.
Lake(s) Sampled/Region
Observation
Reference
Brooktrout Lake, Adirondacks
Fragilariforma acidobiontica, a diatom that is often
abundant at pH <5.0, was present in lake sediments
deposited since the 1950s.
Sutherland et al. (2015)
Glasgow Lake, Nova Scotia
Diatoms inferred to have responded to acidification in
response to a peak S loading of about 14 kg S/ha/yr
during the mid-1970s.
Gerber et al. (2008)
Four sampled lakes in Nova
Scotia
Decline in Daphnia beginning in the early 20th century.
Korosi and Smol (2012)
36 small headwater lakes in
the Boreal Shield of south
central Ontario
Large declines documented in relative abundances of
Ca rich Daphnia spp. and increases in the Ca poor
species Holopedium glacialis.
Jeziorski et al. (2012a)
Lakes downwind of iron
sintering plant northeast of
Wawa, Ontario
Recent cladoceran sedimentary assemblages remain in
an altered state, although lakes have returned to
circumneutral pH.
Jeziorski et al. (2013)
Low-Ca lakes (mean 2 mg/L)
from the Experimental Lakes
area, Ontario
Present-day cladoceran assemblages including
Bosmina spp. and H. glacialis differed from preindustrial
assemblages.
Jeziorski et al. (2014)
Middle Lake (limed in 1973) in
the Experimental Lakes area,
Ontario
Cladoceran assemblages have not returned to
preacidification levels, and many rare species are not
present.
Labai et al. (2014)
Two boreal lakes in the
Killarney lakes region, Ontario
recovering from acidification
Cladoceran assemblages varied little from preindustrial
times to recovery. Bosmina spp. dominated in the lakes.
Labai etal. (2015)
54 lakes in the Muskota-
Haliburton region of Ontario
Numbers of planktonic diatoms such as Cyclotella
stelligera have increased with lake DOC and warming
from 1992 to present.
Hadlev etal. (2013)
Four boreal lakes recovering
from acidification and four
minimally disturbed lakes in
Sweden
Phytoplankton and littoral assemblages in acidified lakes
become more similar to reference lakes overtime.
Differences were greatest for phytoplankton
assemblages.
Johnson and Anqeler
(2010)
Ca = calcium; DOC = dissolved organic carbon; L = liter; mg = milligrams; S = sulfur.
1
2
3
4
June 2018	8-43	DRAFT: Do Not Cite or Quote
In a lake study from Sweden, Johnson and Angclcr (2010) compared phytoplankton and
littoral assemblages in four boreal lakes recovering from acidification and four minimally
disturbed reference lakes over a two-decade period. They observed that assemblages in
the previously acidified lakes became more similar to assemblages in reference lakes over

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
time. Differences were greater for phytoplankton assemblages than for invertebrate
assemblages. Biological responses were correlated with inter-annual variability in
climate, in addition to decreased water acidity. Hadlev et al. (2013) conducted a resurvey
of 54 lakes in the Muskota-Haliburton region of south-central Ontario. They documented
recent changes in chemistry and biology for the lakes using sediment cores and diatom
transfer functions. Observed pH increases were accompanied by decreases in acidophilic
diatoms. However, diatoms were observed to be moving toward a novel assemblage
rather than the preacidification assemblage. Lake DOC concentration has increased from
1992 in the lakes sampled Hadlev et al. (2013). as have the numbers of planktonic
diatoms such as Cyclotella stelligera commonly linked to climate warming. The
concentration of Ca has decreased and land use changes have further impacted the lakes.
All of these factors may be influencing diatom assemblage recovery.
Sixteen lakes in Cape Breton Highlands National Park in Nova Scotia were assessed
through sedimentary records for historical trends in water chemistry and changes in
diatom assemblages. Only one lake (Glasgow Lake) was inferred to have acidified under
peak S deposition (about 438 eq S/ha/yr) in the 1970s (Gerber et al.. 2008). Six of the
lakes, including Glasgow Lake, were modeled using the MAGIC model. The five other
lakes that were modeled showed simulated patterns of acidification and subsequent
chemical recovery but no evidence of acidification from the diatom analysis (Gerber et
al.. 2008). The authors suggested that the diatom approach was conservative and/or that
the MAGIC model slightly overestimated changes in acid-base chemistry.
8.4.2. Zooplankton Recovery
Several studies reviewed in the 2008 ISA reported zooplankton recovery in response to
experimental deacidification of lakes. Zooplankton recovery was observed in
experimentally acidified Lake 223 in the Experimental Lakes region in Canada as pH
increased back to 6.1 (Mallev and Chang. 1994). A lag of 1 to 6 years was observed in
recovery of zooplankton in the experimentally acidified Little Rock Lake in Wisconsin
(Frost et al.. 2006). Locke and Sprules (1994) reported that acidification below pH 5 in
the 1970s overcame the resistance stability of the zooplankton community in Ontario
Precambrian Shield lakes. The subset of study lakes that showed pH recovery from
acidification 20 years later in 1990 also showed recovery in the stability of the
zooplankton community. Holt and Yan (2003) also noted recovery in zooplankton
community composition (based on similarity to neutral lakes) in the subset of Killarney
Park (Ontario) lakes in which the pH increased to over 6 during the 1971 to 2000 study
period. They did not, however, note differences in the increase in species richness
between the recovering lakes and nonrecovering lakes.
June 2018
8-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Studies published since the 2008 ISA provide further evidence for plankton recovery. In
Brooktrout Lake in the Adirondacks, phytoplankton and rotifer taxonomic richness
showed substantial increases (Figure 8-9) in association with pronounced decreases in
lake SOr . H+, and inorganic A1 concentrations (Table 8-6). In contrast, species richness
of crustaceans changed little. Sutherland et al. (2015) suggested that the absence of
crustacean recovery may have been due to the severity of the initial stress and continuing
toxic pH conditions during some seasons, and/or to a large population of predatory
chaoborus larvae present since the early 1980s when fish were absent from the lake.
Paleolimnological evidence of zooplankton responses to changing lake chemistry
published since the 2008 ISA is summarized in Table 8-5.
Evidence from newly published studies in lakes near the former smelting complexes at
Sudbury, Ontario, Canada indicate some evidence of biological recovery, although the
relative recovery due to decreasing acidification versus recovery from decreased metal
emissions and deposition is not clear. Valois etal. (2011) surveyed zooplankton
community structural changes related to gradients of acidification, metal contamination,
trophic status and lake depth in 87 lakes around the smelter area,. At pH >6.0, community
composition of copepods and cladocerans did not differ substantially from those in
reference lakes. Recovery was evident in lakes that had pH >5.5, with decreased
community richness of copepod communities in lakes with lower pH. In the lakes around
Sudbury, the survival of some zooplankton species, especially Daphnia mendotae, is
limited by fish populations, notably yellow perch [Perca flavescens; (Valois et al..
2010)1. The relative importance of changing acidity and metal contamination in driving
the observed biological responses is not known. Nevertheless, results suggest that the
re-establishment of the zooplankton community in lakes recovering from acid and metal
stress may require improvements in both habitat quality and the integrity of the food web.
Babin-Fenske et al. (2012) reported results of phylogenetic analyses of freshwater
amphipod (Hyalella spp.) colonization in lakes near Sudbury. Two major groups of
Hyalella were revealed, based on mitochondrial cytochrome c oxidase I sequences. One
group was associated with recolonization of lakes that had historically been most
acidified. The second group was associated with the more peripheral, less impacted,
lakes.
8-45
DRAFT: Do Not Cite or Quote

-------
160 -|
140
- 120
>•
1100
i
E
2." 80
ar
9/
| 60
Ł 40 -\
20
0
20
18
16
I"
^12
<
J10
O"
i 8
c
Ł 6
4
2
0
40
35
1 30
JC
u
= 25
.2
120
I 15
(a)
8 °
8
R2 = 0.88
(b)
(c)
| 10
o Mid-summer Epilimnetic S042' (peq I"1)
• Wet Deposition S042 (meq/m2/yr)
-¦O... ° § Q n
o o °	o
O o	o
o	° e;
® « a 8
® § 8 s
1 8 e
R2 = 0.78

~ Mid-summer Epilimnetic IMA (pM I1)
¦ Mid-summer Epilimnetic [H*] (peq I *)
R2 = 0.56
D o
D ~
I ¦
R2 = 0.55
¦ a
-s.t
~ .'«!¦ i J ¦ ' I j n
. R2 = 0.60
i '
A A
AAA
* i 4 4 R2 = 0.60
a Phytoplankton
* Total Plankton
84 86 88 90 92 94 96 98 00 02 04 06 08 10 12
Year
H = hydrogen; IMA = inorganic monomeric aluminum; peq = microequivalent; pM = micromolar; m = meter; S042" = sulfate;
yr = year.
Source: Sutherland et al. (2015).
Figure 8-9 (A) Midsummer sulfate concentration in the epilimnion of
Brooktrout Lake (o) and in annual wet deposition (•) at local
National Atmospheric Deposition Program/National Trends
Network Station NY52 from 1984-2012. (B) Midsummer
epilimnetic concentrations of inorganic monomeric aluminum (~)
and [hydrogen ion] (¦) in Brooktrout Lake from 1984-2012.
(C) Midsummer phytoplankton (A) and total plankton
(phytoplankton, rotifers, crustaceans) (A) species richness in
Brooktrout Lake from 1984-2012.
June 2018
8-46
DRAFT: Do Not Cite or Quote

-------
Table 8-6 Midsummer values and productivity analytes in the epilimnion of
Brooktrout Lake in the Adirondack Park from the 1980s through
2010-2012.a
Analyte
(a) 1980s Mean (±SD)
(b) 2010-2012 Mean (±SD)
[H+] (peq/L)
8.15 (2.54)
1.30 (0.26)
ANC (peq/L)
-1.90 (10.31)
11.73 (3.77)
SCM2" (peq/L)
118.18 (9.16)
52.09 (3.88)
NOs" (peq/L)
17.06 (7.36)
7.52 (2.43)
Total P (|jg/L)
3.29 (1.80)
4.00 (2.09)
DOC (mg/L)
0.71 (0.49)
2.40 (0.46)
Ca (peq/L)
68.22 (7.56)
41.65 (12.46)
Secchi (turbidity) (m)
8.89 (1.97)
5.92 (0.66)
Chi a (|jg/L)
0.85 (0.68)
1.86 (1.04)
Reactive silica (mg/L)
3.31 (0.21)
2.39 (0.28)
Total Al (|jM/L)
19.11 (7.13)
7.52 (3.10)
Labile monomeric Al (|jM/L) [IMA]
12.17 (2.93)
1.02 (0.77)
Al = aluminum; ANC = acid neutralizing capacity; Ca = calcium; Chi a = chlorophyll a; DOC = dissolved organic carbon;
H+ = hydrogen ion; IMA = inorganic monomeric aluminum; L = liter; |jeq = microequivalent; |jg = microgram; |jM = micromolar;
m = meter; mg = milligrams; N03" = nitrate; P = phosphorus; SD = standard deviant; S042" = sulfate.
aValues presented in (a) and (b) are means (± standard deviation).The 1980s samples were collected in 1984 (three dates), 1987
(four dates), and 1988 (two dates). The values presented in (b) represent 10 sampling dates during 2010-2012.
Source: Sutherland et al. (2015).
1	In a synthesis of 21 regional surveys of zooplankton recovery in lakes affected by acidic
2	deposition in North America and Europe, Gray and Arnott (2009) identified the most
3	commonly used metrics and factors indicating and limiting recovery of acid-impacted
4	lakes. In the studies evaluated, species richness, the presence of indicator species, and
5	relative species abundances were commonly assessed. At pH >6.0, recovery of
6	zooplankton was significant, although often incomplete in both North American and
7	European lakes. The authors identified slow chemical recovery, dispersal limitation, and
8	community resistance as primary factors limiting biological recovery although the
9	relative importance varied across the lakes and regions evaluated.
June 2018	8-47 DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Gray et al. (2012) assessed the chemical and biological recovery of 45 previously
acidified lakes in Killarney Park, Ontario. Zooplankton data were compared for the
periods 1972-1973 and 2005. Significant increases in pH overtime were documented for
the majority of the study lakes. Species richness and diversity of many acid-damaged
lakes only increased to a minor extent. They observed differing importance of biotic and
abiotic conditions and dispersal processes in determining the relative contribution of
copepod and cladoceran zooplankton in the post-recovery community structure,
suggesting different expectations for recovery among different taxonomic groups. In a
study of zooplankton communities in four of the lakes, dispersal limitation was identified
as a factor impeding biological recovery (Gray and Arnott. 201IV The researchers
observed a relatively quick return of acid-sensitive zooplankton species that disperse
from streams and the egg bank and a slower return of species that depend on overland
dispersal via wind or animal transport.
Several studies have focused on changes in cladoceran assemblages in response to lake
chemistry. Cladocerans require relatively large amounts of available Ca for growth and
survival. Soil Ca pools and drainage water concentration can be reduced under acidifying
conditions. Under conditions of partial chemical recovery of previously acidified surface
waters, Ca concentrations have remained chronically low in some waters in some regions,
likely impeding biological recovery (Jcziorski et al.. 2012a).
An important approach for studying biological effects of aquatic acidification entails
long-term monitoring of biological community composition. This can be done via
evaluation of sedimentary remains of aquatic organisms in sequential slices of lake
sediment cores. Cladocerans have commonly been the focus of long-term monitoring and
paleolimnological studies to document biological responses to changes in water
chemistry and trophic structure (e.g., Cooper et al.. 2016; Nevalainen et al.. 2014;
Davidson et al.. 2010a; Kurek et al.. 201 ObHc.g.. Cooper et al.. 2016; Nevalainen et al..
2014; Davidson et al.. 2010a; Kurek et al.. 20 lObHc.g.. Cooper etal.. 2016; Nevalainen
et al.. 2014; Davidson et al.. 2010a; Kurek et al.. 201 ObHCoopcr et al.. 2016; Nevalainen
et al.. 2014; Davidson et al.. 2010b; Kurek et al.. 2010a).
Jeziorski et al. (2008) documented major reductions in the abundance of Ca-rich Daphnia
spp. in association with decreases in the concentration of Ca in lake water. They reported
that a high proportion of Canadian Shield lakes had Ca concentration near or below the
threshold level (1.5 mg/L) for decreased survival and fecundity in laboratory studies. In a
study from Nova Scotia, cladoceran remains from three of four sampled lakes showed a
decline in Daphnia beginning in the early 20th century that suggested limnological
changes in response to acidic deposition or increased fish predation (Korosi and Smol.
2012). Following closure of a S emitting iron sintering plant in 1998 in the region
8-48
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
northeast of Wawa, Ontario, pH from three previously acidified lakes returned to
circumneutral conditions due to the high buffering capacity of the local geology.
However, biological recovery has lagged behind, with no recovery observed in
cladoceran sedimentary assemblages (Jeziorski et al.. 2013). Analysis of sedimentary
cladoceran assemblages from low-Ca lakes (mean < 2 mg/L) in the Experimental Lakes
area of Ontario, Canada, Jeziorski et al. (2014) indicated differences between present-day
and preindustrial cladoceran assemblages. In Middle Lake, a lake in the Experimental
Lakes area that was limed in 1973, alterations in cladoceran assemblages were assessed
using the sedimentary record (Labai et al.. 2014). Overall, the results suggested that
biological recovery has not occurred in the lake because many rare species have not yet
returned to preacidification levels.
Bosminids dominated the sedimentary cladoceran assemblage from two acidified lakes in
the Killarney Lake region in Ontario, and minimal shifts were observed in the
paleolimnological record from preacidification conditions to recovery (Labai et al..
2015). In contrast, cladoceran abundance in acidified lakes near Sudbury, Ontario
experienced greater shifts in cladoceran abundance best explained by contamination by
copper (Cu) and nickel (Ni). In an earlier study, Jeziorski et al. (2012a) documented large
declines in the relative abundances of Ca rich Daphnia spp. and increases in the Ca poor
species H. glacialis in 36 small headwater lakes in the Boreal Shield of south central
Ontario having Ca concentration <3 mg/L. Reid and Watmough (2016) estimated that
timber harvesting at planned levels in the Muskoka River watershed in Ontario, Canada,
will cause approximately a 30% increase in the number of sampled lakes that decline to
Ca levels below 1 mg/L, too low to support some key aquatic species, including D. pulex,
compared with a scenario of no further tree harvesting.
Vrba et al. (2016) evaluated the recovery of planktonic and littoral invertebrate
communities over 12 years (1999-2011) in eight Bohemian Forest lakes in Europe.
During the decades of the 1990s and 2000s, S and N deposition decreased by 86 and
44%, respectively. Half of the study lakes (those having Al concentrations lower than
200 (ig/L [7.4 |iM |) showed some degree of biological and chemical recovery. The lakes
having Al concentrations higher than 200 (ig/L (7.4 |iM) did not show pronounced
decreases in lake acidity or improved biological condition. Differences between data
collected in 1999 and data collected in 2011 showed pronounced differences in biological
recovery depending on the Al conditions, with largest recovery responses observed for
zooplankton. Such changes were driven by species gains in ciliates and crustaceans and
replacements of species of rotifers in the high-Al lakes. Changes in the low-Al lakes were
dominated by both species gains and losses. Results of multivariate analyses suggested
that the major driver of differences in biological recovery was Al concentration. Species
in low-Al lakes responded to improved chemistry by exhibiting gradual patterns of
8-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
change. Recovery of high-Al lakes was constrained by a toxicity threshold response.
Overall, biological recovery in these Bohemian lakes lagged behind recovery of water
acid-base chemistry. Species gains and losses were due to responses of many aquatic
species rather than to changes in a relative few indicators.
Effects of regional landscapes on trait divergence in a freshwater copepod
(Leptodiaptomus minutus) were investigated by testing differences in acid tolerance in
three spatially distinct groups of lakes in northern Canada: (1) exclusively circumneutral
lakes (pH >6.0) in Quebec, (2) heterogeneous mix of acidic and circumneutral lakes of
Killarney, Ontario, and (3) a mainly acidic group of largely bog-influenced lakes
interspersed with some circumneutral lakes at Cape Race in Newfoundland (Dastis and
Perry. 2016). Acid-tolerant L. minutus from an acidic source pond survived better
subsequent to transfer to both low (3.6) and high (7.0) pH conditions. Nevertheless,
copepod survival was dependent on the population source. Copepods from the
circumneutral source pond were negatively impacted after 5 days of exposure to water
having pH 3.6. For copepods from the acidic source pond, survival was higher and the
negative impact occurred after a much longer period of time, 17 days. The authors
interpreted this result as being indicative of a fitness trade-off regarding tolerance of
acidity. To assess community responses to multiple stressors, the researchers contend that
it will be important to understand contingencies that are landscape-dependent, including
pH tolerance.
8.4.3. Benthic Invertebrate Recovery
In the 2008 ISA, few studies reported benthic recovery in previously acidified waters.
One study by Soulsbv et al. (1995) indicated recovery in some streams in Scotland that
experienced ANC increases although recovery was not observed in the most severely
acidified streams.
Since the publication of the 2008 ISA, additional studies have been conducted that assess
recovery of benthic organisms. Most of that research has been conducted in Canadian and
European waters. Indices to evaluate biological effects of aquatic acidification on benthic
communities commonly rely on species of Ephemeroptera, Plecoptera, Trichoptera,
Gastropoda, and Crustacea. In a study from the Neversink River basin in the Catskill
Mountains in New York, Burns et al. (2008^) collected water chemistry data and
surveyed macroinvertebrates, fish, and periphytic diatoms from 12 streams. This survey
was conducted in 2003 and compared with a 1987 survey conducted at the same
locations. Differences between the two surveys were not significant overall. However, a
shift toward a less acid-tolerant macroinvertebrate community in the 2003 survey in
June 2018
8-50
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
streams in the upper river basin that were the most acidic in 1987 suggests biological
changes that are consistent with improvement in water chemistry. Further monitoring will
be needed to determine the eventual extent of biological recovery.
Crayfish and other benthic invertebrates that require large amounts of Ca for growth may
be reduced or absent in acidified water bodies that have experienced Ca depletion,
reducing Ca bioavailabity, caused by acidification(U.S. EPA. 2008a; Baker etal.. 1990a).
Hadlev etal. (2015) studied the limnological record of four lakes in Algonquin Park,
Ontario where native populations of the crayfish Cambarus bartonii have not recovered
despite improvements in pH. Cladoceran remains were used as proxies for historical Ca
trends. Ca levels in the lakes are currently <2 mg/L, which is below the minimum
requirement for crayfish survival (2 to 10 mg/L). Depletion of Ca in the soils and reduced
Ca export from soil to lakes appear to contribute to the lack of recovery of crayfish
populations in the region.
Two studies examined the response of littoral benthic macroinvertebrate community
composition to changing water chemistry in 17 lakes on the Precambrian Shield in
eastern Canada that were recovering from acidification. The first study Lento et al.
(2008) tracked benthic invertebrate community composition over a 14-year period
(1988-2002). There was a strong correlation between changes in chemical variables and
shifts in benthic invertebrate communities. The rate of change in invertebrate community
composition was greatest between 1993 and 1997, which corresponded most closely with
water chemistry changes. The second study (Lento et al.. 2012) evaluated temporal trends
in 15 years of data from the same lakes. In analyses of both single and multiple lakes,
decreasing proportions of Chironomidae and increasing proportions of Ephemeroptera,
Plecoptera, and Tricoptera (quantified collectively as the EPT Index) were observed. Six
of the nine lakes that showed significant recovery trends in more than one benthos metric
exhibited a significant decrease in Chironomidae and concurrent increase in EPT. The
results of these two studies, thus, suggested that the benthic macroinvertebrate
communities of these study lakes changed in response to deacidifying changes in lake
chemistry by shifting towards more acid-sensitive taxa. The observed response was
consistent with benthic invertebrate recovery from acidification.
Some species of the order Odonata (including dragonflies) respond to changes in water
acidity. For example, reduced water acidity in response to surface water liming can cause
direct or indirect effects on some of the most common dragonfly species. (Al Jawaheri
and Sahlen. 2017) investigated aquatic dragonfly communities in 47 Swedish lakes.
Several were currently (n = 13) or previously (n = 8) limed. Seven of the most common
dragonfly species were less common in limed lakes as compared with reference lakes that
8-51
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
were not limed perhaps due to fish predation. A1 Jawaheri and Sahlen (2017) suggested
use of this order of stream invertebrate as indicators of water quality.
In a bioassessment in the Athabasca oil sands region of Alberta, Canada, Parsons et al.
(2010) sampled 32 lakes to assess potential impacts of pollutant emissions and acid
deposition on benthic macroinvertebrate communities. Five of the lakes were selected
from an area with modeled high deposition. All of the statistical methods used (one
sample /-tests, multivariate analysis of variance, and test site analysis) showed that
assemblages differed between test lakes and reference lakes. However, results suggested
these differences were more due to intrinsic lake physicochemical factors, such as
geology and water chemistry, than to atmospheric deposition.
Several studies conducted in Europe have also considered macroinvertebrate recovery
following ongoing chemical recovery. Frame et al. (2016) tested whether the growth of
Baetis rhodani (an acid-sensitive mayfly nymph) is decreased by competition with
Leuctra inermis (an acid-tolerant stonefly nymph) in a stream in the U.K. recovering
from acidification. This experiment tested the biotic resistance hypothesis, which
suggests that past acidification of surface waters has changed acidified waters to an
extent that prevents reinvasion of acid-sensitive species even if acidity is substantially
reduced. Introduction of Baetis to water containing different densities of Leuctra had no
effect on Baetis growth. In one of the previously most acidified areas of Europe,
Svobodova et al. (2012) compared water chemistry and macrozoobenthos composition of
the outflows from two lakes in the Bohemian Forest. The water chemistry was recovering
in response to reduced levels of acidic deposition. Evidence of biological recovery was
found for Lake Laka (current pH ~5.2). but not for the more strongly acidified Lake
Certovo (pH ~4.6). At the Lake Laka outflow, increasing numbers of Ephemeroptera and
Tricoptera taxa were observed. Comparable changes were not observed at the outflow for
the more strongly acidified lake. In an analysis of a 20-year data set (1988-2008) from
the Acid Waters Monitoring Network (AWMN) in the U.K., macroinvertebrate recovery
was observed in 5 of the 11 stream sites and at 5 of 12 lakes (Murphy et al.. 2014). An
increase in ANC over time was observed at all 10 sites where biological recovery was
evident based on trends in the macroinvertebrate community. An additional seven sites
showed ANC recovery but no recovery of the macroinvertebrate community. The authors
attributed this observation to differences in chemical environments and/or ecological
interactions that interfere with recovery such as food-web dynamics and predation. In
another 20-year data set, Angeler and Johnson (2012) analyzed littoral invertebrates in
acidified and circumneutral lakes in Sweden. The concentration of SOr decreased in
both acidified and circumneutral lakes but converged in the latter part of the study.
Chemical recovery was generally weak in the acidified lakes, with pH increasing by only
about 0.1 to 0.2 pH units in the study lakes. These results from Sweden suggest that
8-52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
because invertebrate communities respond to complex processes, including both local
and regional factors, partial recovery of water chemistry from acidification is at times not
accompanied by observable biological recovery, especially as related to biodiversity.
Data from two national-scale benthic macroinvertebrate surveys, the AWMN in the U.K.
and a lakes survey in Norway, were used to compare field sampling results to predictions
of maximum species richness as a function of chemical conditions in surface waters
recovering from acidification (Stockdale et al.. 201IV In general, there was good
agreement between model predictions and observed trends in EPT taxa. Biological
recovery rates for actual and predicted species richness were generally consistent with
each other, at 1.2 to 2.0 species per decade change. However, actual recovery rates in the
AWMN lakes were less than in the rivers (0.6 vs 2.0 species per decade recovery),
whereas predicted rates were similar (1.7 vs 2.0 species per decade). At sites where there
was poorer agreement between model predictions and observations, differences in water
chemistry explained some of the observed reduction in species richness. The authors
speculated that factors not included in the model, such as biotic interactions and site
conditions, could also affect species richness.
8.4.4. Fish and Amphibian Recovery
Some of the most in-depth studies of the effects of acid stress on fish (Appendix 8.3.6)
have been conducted in streams in Shenandoah National Park, VA (Cosby et al.. 2006)
and lakes in the Adirondack Mountains, NY (Sullivan. 2015). Effects on fish have also
been documented in acid-sensitive streams of the Catskill Mountains of southeastern
New York and the Appalachian Mountains from Pennsylvania to Tennessee and South
Carolina (Baldigo and Lawrence. 2001; Bulger et al.. 2000; Bulger etal.. 1999; SAMAB.
1996; Charles and Christie. 1991). Since the 2008 ISA, several studies have documented
vertebrate recovery that corresponds to decreasing acidic deposition.
Evidence for recovery of fish populations following reduction in acidic deposition or
through liming (Appendix 8.4.6) was reported in several studies reviewed in the 2008
ISA (Gunn et al.. 1988; Kelso and Jeffries. 1988; Beggs and Gunn. 1986; Dillon et al..
1986; Keller and Pitblado. 1986; Raddum et al.. 1986; Hultberg and Andersson. 1982).
There have been additional studies of fish response to surface water acidification and
subsequent recovery since the 2008 ISA. These studies have been conducted in both
North America and Europe.
Sutherland et al. (2015) documented the reintroduction and re-establishment of a
naturalized native fish species (brook trout) in an Adirondack lake (Brooktrout Lake)
from which this species had previously been extirpated. The historical decline in water
June 2018
8-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
quality and loss of the fishery were reconstructed from available data dating back more
than a century. Lake chemistry has improved substantially in response to decreases in
acidifying deposition since the 1980s, with pronounced decreases in midsummer
epilimnetic inorganic A1 and H+ concentrations. In Brooktrout Lake, the mean ANC
increased from -2 j^ieq/L during the 1980s to 12 j^icq/L during the period 2010-2012
(Table 8-6) . Substantial changes were also noted for other variables, including SO42 . H+,
NC>3~, DOC, and inorganic Al. During the period from 2005-2007, brook trout were
reintroduced to the lake and have subsequently survived and successfully reproduced. In
2012, young brook trout were observed and photographed, documenting reproduction in,
or in tributary streams near, the lake. Other acid-impacted lakes in the Adirondacks such
as Honnedaga Lake lost acid-sensitive fish species between 1920 and 1960 while
continuing to support brook trout (Josephson et al.. 2014). By the late 1970s, brook trout
were also considered extirpated from Honnedaga Lake, but persisted in tributary refugia.
By 2000, brook trout had recolonized the lake coincident with reductions in surface-water
SO42 , NO3 and inorganic Al concentrations.
In addition to these studies of recolonization of lakes by brook trout where acid-base
chemical conditions were improving, new reports of shifts in fish communities in
response to acidification have been published since the 2008 ISA. Warren et al. (2008)
used historical records to document a fish community shift in the upper mainstem of
Hubbard Brook in New Hampshire corresponding to chronically acidified conditions in
the 1970s. Records indicated that there were at least three fish species in the 1960s: slimy
sculpin (Cottus cognatus), blacknose dace (Rhinichthys atratulus) and brook trout. Only
brook trout were present in surveys conducted in 2005, 2006, and 2007. In an evaluation
of fish recovery in nearly 5,000 small headwater lakes in Finland where at the end of the
1980s populations were either diminished by acid deposition or extirpated, Rask et al.
(2014) noted recovery of perch (Perca fluviatilis), a relatively acid-tolerant species,
starting in the 1990s. The perch population structure has returned to normal during the
monitoring period, which ended in the mid-2000s. Little or no recovery of the
acid-sensitive roach (Rutilus rutilus) was observed over that time period.
Despite observed reductions in acidic deposition and improvement in water quality of
New York lakes, Baldigo et al. (2016) found no evidence of widespread or substantial
biological recovery of brook trout populations or broader fish communities in the
Adirondack Mountains. The study focused on 43 lakes sampled by the Adirondack Lakes
Survey Corporation during three time periods (1984-1987, 1994-2005, and 2008-2012).
Metrics reflecting fish species richness, abundance of fish species, and abundance of
brook trout did not change significantly over the 28-year period across the group of study
lakes despite a significant average ANC increase and a decrease in inorganic Al over
time. Fish species richness and catch of all fish species per net-night were positively
8-54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
related to lake chemistry reflecting some limited degree of biological recovery. The
authors speculated that additional time may be needed for fish recolonization.
Recovery of fish populations in streams may be affected by habitat considerations as well
as water chemistry. Warren et al. (2010) evaluated the relative influence of stream habitat
and pH on total fish biomass in 16 streams in the northeastern U.S. Physical and chemical
factors evaluated in the study included total pool area, cover, large wood frequency, and
water temperature. Both pool area and pH were closely correlated with prediction of fish
biomass. Physical habitat is likely more important at pH >5.7 (at low flow) and becomes
less important at lower pH when conditions are too acidic to support fish.
Recovery of young brown trout (Salmo trutta) in acidified streams in southern Norway
was assessed by Hcsthagcn et al. (2016) over the period 1987-2010. The density of fish,
both young-of-the-year and older parr, increased significantly during the study period.
Water quality also improved (pH 5.0-5.5 increasing to 5.8-6.0). Nevertheless, the
densities of both fish age groups (young-of-the-year and older parr) decreased for a
period of time in the early 1990s. This was attributed to seasalt acidification associated
with severe weather conditions that caused increased marine salt deposition. This finding
emphasizes the fact that recovery from acidification should be evaluated in the context of
other changes, including natural or other human-caused disturbance regimes.
Even if the preindustrial acid-base chemistry of a water body is fully restored, there is no
guarantee that any aquatic species that had previously been eliminated in response to
acidification will in fact return to that water body. There may be physical, chemical, or
biological barriers to species reintroduction. Methods have not been well developed and
tested for reintroduction of species other than fish.
Potential impacts of aquatic acidification on amphibian species occur in the context of
multiple stressors, including fish introductions and the presence of fungal pathogens. A
dramatic example of environmental stress in amphibians is the widespread prevalence of
the disease chytridiomycosis, which is caused by the fungus Batrachochytrium
dendrobatidis. Major die-offs of frogs, toads, salamanders, and other amphibians have
been documented in the U.S. and elsewhere (Martel et al.. 2014; Rosenblum et al.. 2010).
These die-offs have not been attributed to acidification.
8.4.5. Bird Recovery
Common loon (Gavia immer) breeding success was assessed from 1982 to 2007 in
38 lakes historically impacted by deposition from smelter activities in the Sudbury,
Ontario region (Alvo. 2009). No fledglings were observed on any lakes that had pH <4.4.
June 2018
8-55
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
In two of the lakes that were previously too acidic for loon reproduction, chicks were
observed as acid conditions improved over the duration of the study. However, it was not
clear from the results whether there is a critical pH for loon fledging survival.
8.4.6. Mitigation
Application of lime or other sources of base cation(s) is well known to increase the
bioavailability of nutrient bases to terrestrial and aquatic biota. Neutralization of acidity
by addition of limestone or other Ca source to watershed soils or directly to water bodies
has been used for several decades in Europe and to a lesser extent in North America to
mitigate effects of acidifying deposition. Analyses of long-term data suggest that liming
watershed soils increases Ca availability to aquatic and terrestrial biota and limits
mobilization of A1 from soils to surface water. Watershed liming of soils generally
promotes long-lasting effects whereas direct liming to streams and lakes causes
short-term chemical fluctuations and less pronounced ecosystem recovery (Lawrence et
al.. 2016). Prior to the 2008 ISA, relatively few studies were conducted in the U.S. that
focused on mitigation of harm to aquatic organisms caused by acidification. Limited lake
liming had occurred in the Adirondacks in an effort to reverse the adverse effects of
water acidification (Schofield andKeleher. 1996).
It has been well documented that the intensity and duration of acidity mitigation achieved
by Ca addition depends to a large degree on the method of Ca application. In particular,
different results have been achieved depending on the site of Ca addition: directly to the
water body and/or more broadly across the watershed that contributes drainage water
(and associated base cations) to the water body in question. Application of lime to the
terrestrial watershed leads to acidity mitigation that is more gradual and of longer
duration than lime application directly to surface waters (Davis and Goldstein. 1988).
A liming study at the Woods Lake watershed in the Adirondack Mountains applied
calcite (CaCCh) to two paired watersheds, with and without pronounced wetland
influence (Driscoll et al.. 1996). Lake water showed gradual improvements in pH, ANC,
and the concentration of Ca at the site of watershed liming, in contrast to the more abrupt
and short-lived effects of direct surface water lime application. The liming decreased the
concentrations of inorganic Al and increased DOC. Water quality improvements affected
both the treated stream and the downstream lake (Burns. 1996). Brook trout spawning
was restored to the tributary stream (Schofield andKeleher. 1996). Only limited
additional research has been conducted in the U.S. in more recent years. Knoepp et al.
(2016) estimated that about 11.6 to 21.1 Mg/ha of CaC03 would be needed to increase
surface soil (upper 30 cm) pH to 5.0 in three high-elevation watersheds in the
June 2018
8-56
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Appalachian Mountains. Stream acid-base chemistry was related to the concentrations of
N and A1 in the soil O-horizon and to the total amount of C and Ca in the soil.
Lawrence et al. (2016) re-examined the liming approach as a method for accelerating the
recovery of acidified ecosystems where chemical and/or biological recovery have lagged
behind reductions in atmospheric S and N deposition. They presented evidence
suggesting that recovery rates are fastest for lakes and slowest for soils, with intermediate
stream recovery rates. They also emphasized that lime should not be applied to
environments, including wetlands, where indigenous species are adapted to naturally
acidic conditions.
An experimental stream and watershed liming study is currently in progress in the
Adirondacks. This work is being conducted by a group of researchers under the
coordination of the U.S. Geological Survey in Troy, NY. Results are not yet available at
the time of this writing.
In an effort to mitigate Atlantic salmon population declines in Norway due to
acidification, liming was implemented in 21 acid-impacted rivers. Hesthagen et al. (2011)
electrofished 13 rivers 1 year before and annually for 12 years after initiation of liming.
Increases in salmon parr densities were slow, and the authors estimated that it will take
more than 20 years of liming to restore lost salmon stocks. Rivers that were stocked and
those lacking hydropower developments generally had higher fry densities and faster
increase in parr densities subsequent to lime application.
Mant et al. (2013) conducted a review and meta-analysis of the impacts of liming streams
and rivers on the biological recovery of fish and aquatic invertebrates. The meta-analysis
included studies in the U.K., Norway, Sweden, U.S., and Canada. On average, liming
increased the abundance and richness of acid-sensitive invertebrates and increased fish
abundance. Nevertheless, benefits were variable and did not occur in all rivers.
The large national-scale lake and stream liming program in Sweden offers a unique
opportunity to evaluate the expected recovery of fisheries in response to water
deacidification at the national level. (Holmgren et al.. 2016) reported results from
monitoring since the 1980s of 1,029 limed streams and 750 limed lakes, plus reference
(unlimed) sites (195 streams and 101 lakes). Overtime, the proportion of limed streams
that had no fish decreased. Species richness and the percentage of streams in which
brown trout (Salmo trutta) was detected both increased over time after initiation of
liming. The abundance of several species, including brown trout, perch (Perca fluviatilis)
and roach (Rutilus rutilus), increased more at sites that were limed as compared with sites
that were not limed. Species richness increased slightly in limed streams (less than 2 on
average before liming to more than 2.6 species 16 years after initiation of liming).
8-57
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Nonlimed streams did not show a change in richness over time. Patterns were less clear
for lakes, perhaps due to the influence of lake size and/or productivity, either of which
might mask the effects of liming.
8.5. Levels of Deposition at Which Effects Are Manifested
The level of N and S deposition and their impacts vary across the landscape. Not all
environments are sensitive to acidifying deposition at the levels seen in the past and
present. Effects on biological communities are related to both the characteristics
(Appendix 8.5.1) and extent and distribution (Appendix 8.5.2) of acid sensitive
freshwaters across the U.S. as well as climate factors (Appendix 8.5.3). Appendix 8.5.4
includes both empirical and modeled thresholds (critical loads) of effects of acidifying
deposition.
8.5.1. Most Sensitive and Most Affected Ecosystems and Regions
The characteristics of acid-sensitive ecosystems were well known and summarized in the
2008 ISA. In brief, the effects of acid deposition on aquatic systems depend largely upon
the ability of an ecosystem to neutralize additional acid inputs. No one level of deposition
can be used to generalize the sensitivity or impacts of acidifying deposition across a
region or the country. The principal factor governing the sensitivity of aquatic
ecosystems to acidification from acidifying deposition is geology [particularly surficial
geology; (G reaver et al.. 2012)1. Geologic formations having low base cation supply, due
mainly to low soil and bedrock weathering, generally underlie the watersheds of
acid-sensitive lakes and streams. Till thickness has been shown to be a key control on the
pH and ANC of Adirondack lakes (Driscoll et al.. 2016). whereby lakes in watersheds
inferred to be underlain by thin till tend to be highly sensitive to acidification (Driscoll
and Newton. 1985). Figure 8-10 is a map of aquatic ecosystem sensitivity in the eastern
U.S. based on underlying geology in unglaciated areas and ANC in the glaciated region.
Other factors contribute to the sensitivity of surface waters to acidifying deposition,
including topography, soil chemistry and physical properties, land use and history, and
hydrologic flowpath.
June 2018
8-58
DRAFT: Do Not Cite or Quote

-------
Source: Lovett et al. (2009)
Figure 8-10 Map of landscape sensitivity to acidic deposition for the
northeastern and mid-Atlantic U.S. Stippled areas were not
considered.
1
2	The 2008 ISA summarized findings from several national surveys on the effects of
3	acidifying deposition, including the National Lakes Survey and the National Streams
4	Survey conducted in the mid-1980s, the Wadeable Streams Assessment (WSA) in 2004,
5	the U.S. EPA Long-Term Monitoring program that began in 1983, and the Temporally
6	Integrated Monitoring of Ecosystems probability surveys that began in 1991. Results of
June 2018
8-59
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
these surveys suggested that acidifying deposition had acidified surface waters in the
southwestern Adirondacks, New England uplands, eastern portion of the upper Midwest,
forested Mid-Atlantic highlands, and Mid-Atlantic coastal plain (U.S. EPA. 2008a).
At ANC of 100 (j,eq/L or less, acid-sensitive aquatic species may be adversely affected
[Figure 8-4; (U.S. EPA. 2008a; Sullivan et al.. 2006a; Bulger et al.. 2000; Bulger et al..
1999)1. Sensitive water bodies, therefore, can be defined as those that have ANC of
100 |icq/L or less. Sensitivity increases with further decreases in ANC. In general,
streams in the eastern U.S. that are sensitive tend to be headwater streams of first- to
third-order. Streams and lakes in the western U.S. tend to be headwater systems in high
elevation areas. Recent research has not changed this understanding but rather has served
to strengthen it (Appendix 7).
Controls on surface water acidification and recovery are not necessarily spatially
homogeneous. For example, the adsorption and desorption of S on soils can exert major
controls on acidification, and likely chemical recovery from acidification, in the
unglaciated soils of the southeastern U.S. (Rice et al.. 2014). Limitations in the extent of
S driven acidification and recovery attributable to S adsorption is less pronounced in the
northeastern U.S.
8.5.2. Extent and Distribution of Sensitive Ecosystems/Regions
The 2008 ISA summarized the extent and distribution of freshwater ecosystems sensitive
to acidifying deposition. Measured data on lake and stream ANC across the U.S. exhibit
clear spatial patterns. In the U.S., surface waters that are most sensitive to acidification
based on ANC are largely found in the Northeast, southern Appalachian Mountains,
Florida, the upper Midwest, and the mountainous West (Sullivan. 2017; McDonnell et al..
2014b; Greaver et al.. 2012; Campbell et al.. 2004a; Driscoll et al.. 2001b; Baker et al..
1990b; Omernik and Powers. 1983). Figure 8-11 is a national map of surface ANC for
the U.S. that includes nearly 200,000 measurements taken at nearly 20,000 spatially
unique locations sampled between 1980 and 2011 (Sullivan. 2017). Samples expected to
be strongly influenced by acid mine drainage, sea salt spray, or road salt application were
excluded. Thus, the included data represent sites not likely confounded by major
disturbances not related to acidic deposition. Sullivan (2017) found 6,065 sites that had
ANC <100 (ieq/L. Acidic waters were mostly restricted to northern New York, New
England, the Appalachian Mountain chain, upper Midwest, and Florida (see Figure 8-11).
Low, but positive, ANC values were depicted for these same regions plus high-elevation
portions of the West and parts of Arkansas and the Gulf states. These geographical
patterns are thought to largely reflect base cation supply in soils. Levels of acidifying
June 2018
8-60
DRAFT: Do Not Cite or Quote

-------
1	deposition in the West are low in most areas, acidic surface waters are rare, and the extent
2	of chronic surface water acidification that has occurred to date has been limited (Sullivan.
3	2017; Charles and Christie. 1991). Episodic acidification, however, occurs in both the
4	East and West at some acid-sensitive locations. These areas can be identified by CL maps
5	for the U.S. based, for example, on an ANC target of 50 (ieq/L I Figure 8-12; (Blctt et al..
6	2014)1.
June 2018
8-61
DRAFT: Do Not Cite or Quote

-------
Source: Sullivan 12017).
Figure 8-11. Surface water Acid Neutralizing Capacity (ANC) map, based on data compiled by Sullivan (2017)
June 2018
8-62
DRAFT: Do Not Cite or Quote

-------
Legend
Critical Load (N»S eq ha"1 yr*'|
1 -2QC0
MOOt-BOM
¦¦ KOI 12OT0
12001 - :•»»
IB 20001 -41Stt
Smi«
Ugend
Critter Load (N*S eq ha ' yr ^
1-21X0
3001 - 6000
§¦1 6001 • 13000
130D1 -MO®
¦¦ 30001 -41906
Slata
" V 'J
> . •
• f
ha = hectare; eq = equivalent; km = kilometer; L = liter; N = nitrogen; S = sulfur; |jeq = microequivalent; yr = year.
Source: Blett et al. (2014).
Figure 8-12 (a) Minimum critical loads of surface water acidity for nitrogen
and sulfur. Grids represent the minimum calculated critical load
from all data within the 36 * 36-km grid cell (b) Mean critical loads
of surface water acidity. Grids represent the average calculated
critical load from all data within the 36 x 36-km grid cell. The
critical chemical criterion used was an acid neutralizing capacity
50 peq/L.
June 2018
8-63
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
8.5.3.
Climate Modification of Ecosystem Response to N and S
Acidification and recovery of freshwaters will be affected by the physical, chemical, and
biological modifications to acid inputs projected to occur with changes in annual mean
temperature and magnitude of precipitation (Greaver et al.. 2016). Projected shifts in
runoff and timing and quantity of flushing will alter the frequency and duration of
episodic events (Whitehead et al.. 2009). In acid sensitive regions, altered hydrologic
regimes are likely to affect weathering rate of base cations, lake water levels and organic
matter inputs to catchments (Adrian et al.. 2009; Porcal et al.. 2009). Air temperature
increases will lead to warmer surface waters altering thermal stratification of water
bodies, distribution of aquatic biota in streams and lakes and community composition
(Adrian et al.. 2009; Keller. 2007). Climate change may play a role in shifting baselines
for ecosystem recovery in previously acidified lakes (Appendix 8.4). Warren et al. (2017)
suggested that the major stressor affecting native coldwater fish species in the eastern
U.S. is shifting from acidification to thermal stress and some lakes recovering from
acidification may provide a degree of protection against climate affects. As DOC in the
water increases with increasing lake pH in recovering lakes, decreased water clarity may
create cooler refuge habitat for fish. Appendix 13 includes a more detailed discussion of
how climate (e.g., temperature and precipitation) modifies ecosystem response to
acidification.
8.5.4. Critical Loads
A critical load (CL) is a quantitative estimate of exposure to one or more pollutants
below which significant harmful effects on specified sensitive elements of the
environment do not occur according to present knowledge KSprangcr et al.. 2004;
Nilsson and Grennfelt. 1988) Chapter 1.2.2.31. The following sections express deposition
in eq/ha/yr of S, N, or S + N because one or both pollutants could be contributing to the
observed effects.
8.5.4.1. U.S. Critical Loads
The 2008 ISA documented use of the CL approach to determine sensitivity to
acidification. Critical load applications for surface water in the U.S. have been reviewed
by Porter et al. (2005). Burns et al. (2008a). Pardo et al. (2011a). NAPAP (2011). Lovett
(2013). and Sullivan and Jenkins (2014). Most aquatic CL studies conducted in the U.S.
have used surface water ANC as the metric of water quality change in response to
acidifying deposition. This should not be taken as an indication that ANC should
June 2018
8-64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
necessarily be the only environmental predictor of harm to aquatic species on which to
base CLs. Other potentially useful variables include water pH or BCS. Either, or both,
can be used instead of, or in addition to, ANC.
The initial step in developing a CL is to define a threshold for a significantly harmful
effect. This biological change can be defined for the protection of a single species or an
entire biological community, and defining varying levels of protection for either can be
achieved through adoption of different indicators of harm. Because CLs depend on what
is being protected, any given biological community requires a set of CLs. The closely
related target load (TL) estimates the reduction needed to achieve a certain condition at a
designated future time period. A target load can be less than, equal to, or greater than the
CL. As described in the 2008 ISA, there are different types of CLs: empirical CL,
modeled steady-state CL (time invariant), and modeled dynamic CL [change through
time with a specific target date, often to the Year 2100 as a management TL; (U.S. EPA.
2008a)]. Several CL studies were summarized in the 2008 ISA.
As discussed in the 2008 ISA, a useful aspect of the CL approach is the calculation of
exceedance. Knowledge of where and to what extent ambient air pollutant loads exceed
levels that are sustainable without causing ecological harm can inform vulnerability
assessments (Sullivan and Jenkins. 2014). Therefore, the exceedance is calculated as the
difference between the ambient deposition of S, N, or S + N and the CL or the TL. It can
be expressed as an absolute deposition amount or as a percentage.
Since 2006, the primary forum for coordination of research on CL development in the
U.S. has been the Critical Loads of Atmospheric Deposition (CLAD) science committee
of the National Atmospheric Deposition Program [NADP; (Blett et al.. 2014)1. Efforts are
underway in association with CLAD to collect needed data and improve methods for CL
calculations throughout the U.S.
Empirical Critical Loads
An empirical CL is developed from observational spatial or temporal gradient studies or
additions of pollutants to determine the deposition load at which chemical or biological
changes (e.g., changes to foliage, lichens, soil, aquatic chemistry/biota) occur in the
environment. Empirical CLs are generally applied to sites or landscapes that are
ecologically comparable to location(s) from which the CLs were determined (cf. Pardo et
al.. 201 la). Empirical CLs for acidification are difficult to apply to sites other than those
used in defining them and to generalize across the landscape because sensitivity to
acidification is highly spatially variable (see Appendix 8.5. IV A given CL for
8-65
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
acidification often only applies to a narrow subset of environmental conditions, in some
cases only to a single water body.
At the time of the 2008 ISA, very few empirical CLs were available for U.S. freshwater
ecosystems. Rocky Mountain National Park has been the site of research addressing the
environmental effects of N deposition. Williams and Tonnessen (2000) reported ANC in
surface water <0 j^icq/L near the park as a result of acidifying inputs of N deposition,
suggesting that current deposition levels are having an observable impact on catchments
in the Colorado Front Range. The authors recommended a CL of 286 eq N/ha/yr to
prevent episodic freshwater acidification in alpine lakes (ANC < 0 (ieq/L). However,
determination of biological change in the field was not a focus of this study.
Other studies conducted in the western U.S. have identified CLs for freshwater systems
(Table 8-7). In Moat Lake in the Sierra Nevada mountains, acidic deposition
(S042 + NO3 ) equal to about 74 eq/ha/yr was correlated with the decline in ANC
observed in the lake between 1920 and 1930 (Heard ct al.. 2014). This was taken by the
authors to be the CL to protect against acidification for Moat Lake, but they did not
specify the level of acidification or how the biological community would be protected.
Diatom-reconstructed ANC of Moat Lake changed from near 100 j^icq/L prior to 1920 to
near 60 (ieq/L during the 1970s. Reconstructed ANC patterns were not correlated with
climate, productivity, or NOx emissions.
Although most CLs are focused on chronic acidification, and by strict definition as a
steady-state metric must be, Baron et al. (2011b) estimated CLs to be about
571 eq N/ha/yr in the Northeast and 286 eq N/ha/yr in the Rocky Mountains for NO3
concentrations as triggers of episodic acidification. CLs for N deposition in California
were estimated based in part on changes in NO3 leaching in stream water, which can
cause or contribute to water acidification (Fenn et al.. 2008). The empirical CL and the
CL simulated by the DayCent model for NO3 leaching were both 1,214 eq N/ha/yr.
Nitrogen deposition exceeds that level at some locations in California.
Modeled Critical Loads
Steady-state and dynamic models are used to quantify relationships between deposition
and biogeochemistry for watersheds in order to develop CL estimates (see description of
models in Appendix 7). Modeled CLs for aquatic acidification are summarized in
Table 8-8.
8-66
DRAFT: Do Not Cite or Quote

-------
Table 8-7 Recent empirical critical loads to protect against aquatic acidification
in U.S. ecosystems.
Ecosystem
Site
Critical Load for N
Deposition
Response
Study
Alpine lakes
Central
Rockies/Colorado
Front Range
286 eq N/ha/yr
Episodic freshwater
acidification
Williams and
Tonnessen(2000)
Mediterranean
California stream
water
California
1,214 eq N/ha/yr
Changes in NO3"
leaching in stream
water
Fenn et al. (2008)
Hardwood and
coniferous
forests drainage
water or stream
Northern forests
571 eq N/ha/yr
Increased surface
water NO3"
leaching
Pardo et al. (2011a)
Aber et al. (2003)
Eastern
hardwood forests
drainage water
Eastern temperate
forests
571 eq N/ha/yr
Increased surface
water NO3"
leaching
Pardo et al. (2011a)
Aber et al. (2003)
Mixed conifer
forests drainage
water
Mediterranean
California
1,214-1,850 eq N/ha/yr
Increased surface
water NO3"
leaching
Pardo et al. (2011a)
Fenn et al. (2010) Fenn
et al. (2008) Breiner et
al. (2007)
Pine forest
drainage water
Temperate Sierra
Nevada
1,071 eq N/ha/yr
Elevated NO3" in
spring and stream
water
Fenn and Geiser
(2011) Pardo et al.
(2011a)
N rich forests
drainage water
Subtropical humid
forests
<357-714 eq N/ha/yr
NO3" leaching
Pardo et al. (2011a)
Lakes
Western U.S.
143 eq N/ha/yr
NO3" leaching
Pardo et al. (2011a)
Baron (2006)
Lakes
Eastern U.S.
571 eq N/ha/yr
NO3" leaching
Pardo et al. (2011a)
Aber et al. (2003)
High elevation
lakes
West and Northeast
571 eq N/ha/yr
(Northeast)
286 eq N/ha/yr (West)
Episodic freshwater
acidification
Baron et al. (2011b)
High elevation
lakes
Sierra Nevada
74 eq/ha/yr
Lake acidification
as measured by
change in ANC
Heard et al. (2014)
ANC = acid neutralizing capacity; eq = equivalent; ha = hectare; kg = kilograms N = nitrogen; N03- = nitrate; yr = year.
June 2018
8-67
DRAFT: Do Not Cite or Quote

-------
Table 8-8 Recent aquatic critical load and target load modeling studies in the
U.S. to protect against aquatic acidification.


Type of



Region
Model
Ecosystem
Focus
Critical Loads
Publication
Conterminous
SSWC and
Lakes and
Implementation
A consistent CL process was
Blett etal. (2014)
U.S.
empirical
streams
of a consistent
documented.




process for





calculating and





mapping





steady-state





CLs


Adirondack
MAGIC and
Lakes
TL for lakes in
To achieve ANC = 50 peq/L
Sullivan et al.
Mountains
regional

2050 and 2100
in 2100, about 30% of lakes
(2012a)

extrapolation


had simulated TL of S


of model


deposition <500 eq/ha/yr and





about 600 lakes were in





exceedance.

Adirondack
PnET-BGC
Lakes
TL link to fish
The magnitude of simulated
Zhou etal. (2015b)
Mountains


and zooplankton
historical acidification




richness
represented by ANC loss





ranged from about 26 to

100 peq/L. The amount of
historical acidification of the
lakes was related to the total
ambient deposition of
SO42" + NO3-, the Ca
weathering rate, and the
simulated preindustrial ANC
in the Year 1850. Adirondack
lakes have lost fish and total
zooplankton species richness
beginning with the onset of
acidic deposition. Modeling
results suggested that
complete recovery to
preindustrial conditions may
not be possible for most
acidified lake ecosystems.
June 2018
8-68
DRAFT: Do Not Cite or Quote

-------
Table 8-8 (Continued): Recent aquatic critical load and target load modeling
studies in the U.S. to protect against aquatic acidification.
Type of
Region	Model Ecosystem Focus	Critical Loads	Publication
Adirondack
PnET-BGC Lake
Effects of
Model simulations suaaested Zhou et al. (2015c)
Mountains

biophysical
that future decreases in SO42"


factors on the
deposition would be more


TL
effective in increasing the lake



water ANC than equivalent



decreases in NO3" deposition.



The difference was a factor of



4.6 during the period 2040 to



2050, but decreased to a



factor of 2 by the Year 2200.



Lakes that had longer



hydrologic residence



exhibited less historical



acidification and therefore



could achieve a higher ANC



in response to chemical



recovery compared with lakes



that had short hydrologic



residence times.
Adirondack PnET-BGC Lakes	TMDLs for	Model simulations suggested Fakhraei et al
Mountains	128 acid-	that an S TL equal to 79 eq (2014)
impaired lakes S/ha/yr (representing a 60%
decrease from ambient
deposition) would lead to
ANC recovery at a rate of
0.18 [jeq/L/yr through 2050,
with reduced rate of recovery
thereafter.
Adirondack
SSWC
Lakes and
Combined
To achieve ANC = 50 peq/L
NAPAP (2011)
Mountain lakes

streams
deposition load
on average, critical load of

and


of sulfur and
sulfur and nitrogen for lakes in

Appalachian


nitrogen to
the Adirondack Mountains is

Mountain


which a stream
1,620 eq/ha/yr, while for

streams


and its
central Appalachian streams




watershed could
is 3,700 eq/ha/yr.




be subjected





and still have a





surface water





concentration





ANC of





50 peq/L on an





annual basis


NY
MAGIC and
Streams
Development
To achieve ANC values of 50
Sullivan (2015)

regional
and lakes
and application
and 20 peq/L in the


extra-

of tools to
Year 2050 and 2100, the TL


polation of

document and
to protect against acidification


model

quantify TL and
of surface waters was




their
exceeded throughout the




exceedances
Adirondack Mountains.

June 2018
8-69
DRAFT: Do Not Cite or Quote

-------
Table 8-8 (Continued): Recent aquatic critical load and target load modeling
studies in the U.S. to protect against aquatic acidification.
Type of
Region	Model Ecosystem Focus	Critical Loads	Publication
Hubbard Brook PnET-BGC Stream
Importance of
Authors developed
Wu and Driscoll
Experimental
incorporating
three-dimensional dynamic TL
(2010)
Forest, NH
base cation
surfaces as function of NO3",


deposition and
S, and base cation deposition


climate in
under varying climate


calculating TLs
scenarios.

Virginia, West SSWC Streams
Estimation of CL
To achieve ANC = 50 peq/L,
Sullivan et al.
Virginia
and exceedance
one-third of the stream length
(2012b)

for stream
in the Blue Ridge ecoregion


watersheds
had CL <500 eq/ha/yr. About


exposed to S
half of the stream reach in the


deposition
study region was in



exceedance under assumed



N saturation at steady state.

Southern SSWC Streams
Regional CLs
To achieve ANC = 50 peq/L,
McDonnell et al.
Appalachian

nearly one-third of the stream
(2014b)
Mountains

length in the southern



Appalachian Mountain region



had a critical load of S

deposition <500 eq/ha/yr,
which was less than the
estimated regional average S
deposition (600 eq/ha/yr).
Due to the local geology,
elevation, and cool and moist
forest conditions, the
percentage of streams in
exceedance was highest for
mountain wilderness areas
and national parks, and
lowest for privately owned
valley bottom lands.
Great Smoky PnET-BGC Streams
TL values for
To achieve ANC values of 0. Zhou et al. (2015a)
Mountains
12 streams to
20, and 50 peq/L, the level of
National Park
achieve ANC of
NO3" + SO42" deposition

0, 20, and
necessary to achieve a given

50 peq/L by
ANC target was

2050
approximately a linear
function of current ANC. Most
streams could not achieve
ANC = 50 peq/L; some could
not achieve ANC = 20 peq/L.
Simulated mean projected
stream ANC of 71 peq/L
(range 32 to 107 peq/L) prior
to industrial development
(-1850) decreases in
response to historical acidic
deposition to 33 peq/L (-13 to
88 peq/L) in 2007.
June 2018
8-70
DRAFT: Do Not Cite or Quote

-------
Table 8-8 (Continued): Recent aquatic critical load and target load modeling
studies in the U.S. to protect against aquatic acidification.
Region
Model
Type of
Ecosystem
Focus
Critical Loads
Publication
Great Smoky PnET-BGC
Mountains NP
Streams TL to achieve The median value of the
preindustrial simulated background (1,850)
ANC and	stream ANC was 50 peq/L,
estimated	with estimates for individual
preindustrial watersheds ranging from 28
ANC minus to 80 peq/L. Simulated ANC
20 peq/L	recovery per equivalent
decrease in the deposition of
NH4+ was more pronounced
than that driven by
comparable decreases in the
deposition of SO42". This
finding was attributed to
continued S adsorption and
low levels of N retention in the
modeled watersheds.
Fakhraei et al.
(2016)
Great Smoky
PnET-BGC Streams
TL to achieve a
Critical loads ranged between
Fakhraei et al.
Mountains NP

pH of 6.0 by
2150
240 and 960 eq/ha/yr of
S042" + NO3-+ NH4+
deposition to eight of the
twelve watersheds. For the
remaining four watersheds,
no reduction in deposition
was sufficient to achieve pH
of 6 by 2150.
(2017a)
Shenandoah
MAGIC Stream
TL values for
To achieve ANC = 50 peq/L
Sullivan et al.


14 streams to
in the year 2100, median
(2008)


achieve
modeled streams located on



ANC = 50 peq/L
siliciclastic geology had TL



in 2100
about 214 eq S/ha/yr,
substantially lower than
ambient deposition.

NE, Rocky
Empirical Lakes
CL of N for
To achieve ANC = 0 peq/L,
Baron et al. (2011b)
Mountains,

acidification of
critical loads to protect

Sierra, NV

lakes
against episodic acidification
were 286 eq N/ha/yr in west
and 571 eq N/ha/yr in NE.

Sierra, NV
SSWC
Lakes
CL of acidity for
208 lakes
To achieve ANC = 10 peq/L,
median CL was 149 eq/ha/yr
and 16-17% of study lakes
were in exceedance.
Shaw et al. (2014)
June 2018
8-71
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
Table 8-8 (Continued): Recent aquatic critical load and target load modeling
studies in the U.S. to protect against aquatic acidification.
Region
Model
Type of
Ecosystem
Focus
Critical Loads
Publication
N/A
Various
Various
Application of
CL and ES
principles for
public land
management
and natural
resources policy
decision making
A conceptual framework was
proposed that illustrates how
CL and ES can be combined,
using as an example aquatic
acidification.
Sullivan (2012)
ANC = acid neutralizing capacity; Ca = calcium; CL = critical load; eq = equivalent; ES = ecosystem service; ha = hectare;
kg = kilogram; L = liter; m = meter; |jeq = microequivalent; MAGIC = Model of Acidification of Groundwater in Catchments;
N = nitrogen; NE = northeast; N03" = nitrate; PnET-BGC = Photosynthesis and Evapotranspiration-Biogeochemical; S = sulfur;
S042" = sulfate; SSWC = Steady-State Water Chemistry; TL = target load; yr = year.
8.5.4.1.2.1. Steady-State Critical Load Modeling
Steady-state CLs can be derived from mathematical mass-balance models under assumed
or modeled equilibrium conditions based in part on water quality measurements. The
models used to derive steady-state CLs vary in complexity with regard to process
representation. However, a fundamental aspect of the various modeling approaches is the
calculation of elemental mass balances.
NAPAP (2011) calculated steady-state aquatic CLs to protect southern Appalachian
Mountain streams and Adirondack Mountain lakes against the combined load of S and N
deposition below which the ANC level would be expected to support healthy aquatic
ecosystems. Research has shown that surface waters with ANC values greater than
50 |icq/L tend to protect most fish (including native [to eastern U.S.] brook trout) and
other aquatic organisms [see Table 8-3. which describes these changes; (Driscoll ct al..
2001b)]. The CL can be calculated to represent the individual or combined deposition
load of S and/or N to which a stream and its watershed could be subjected and still have a
surface water ANC of 50 (j,eq/L on an annual basis.
Sullivan et al. (2012b) and McDonnell et al. (2012) developed an approach for deriving
regional estimates of base cation weathering to support steady-state CL estimates for the
protection of southern Appalachian Mountain streams against acidification in three
ecoregions of Virginia and West Virginia. Weathering estimates at the study watersheds
were developed using MAGIC and extrapolated to the full study region using landscape
characteristics available as regional coverages. Calculated CL values were low at many
locations, suggesting high acidification sensitivity. In the Blue Ridge ecoregion,
calculated CL values to maintain stream ANC at 50 j^ieq/L were less than 500 eq/ha/yr at
June 2018
8-72
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
one-third of the study sites. About half or more of the stream length in the study region
was in exceedance of the CL of S for protecting aquatic resources to an ANC level of
50 (ieq/L over the long term.
In another model simulation for Appalachian Mountain streams, McDonnell et al.
(2014b) calculated critical values, including steady-state aquatic CLs to protect streams
against acidification. They considered an ANC threshold of 50-100 j^ieq/L to be generally
protective of ecological health (cf, U.S. EPA. 2009c; Cosby et al.. 2006). The study area
included mainly streams in southern Pennsylvania, Maryland, Virginia, West Virginia,
North Carolina, Tennessee, and northern Georgia. Weathering values were first
determined using the MAGIC model (Povak et al.. 2014) and extrapolated to the full
study region using machine learning/linear regression models. These values were then
used as inputs to the Steady-State Water Chemistry (SSWC) model to calculate regional
CLs. Nearly one-third of the stream length assessed in the study region had a CL of S
deposition <500 eq/ha/yr, which was below the estimated regional average S deposition
(600 eq/ha/yr). The percentage of streams in exceedance was highest for mountain
wilderness areas and national parks and lowest for privately owned valley bottom lands.
Shaw et al. (2014) applied two variants of the SSWC model (one based on the F-factor
approach and the other determined empirically) to estimate CLs for 2008 lakes in Class I
and II wilderness areas in the Sierra Nevada. It was estimated that slightly more than
one-third of the lakes received acidic deposition higher than the estimated CL. Lakes
included in the model runs were generally dilute, with mean ANC = 56.8 (ieq/L. Critical
loads for acid deposition were estimated to protect ANC to benchmark values of 0, 5, 10,
and 20 (j,eq/L, which span the range of minimum ANC values observed in Sierra Nevada
lakes. Median CLs were 217, 186, 157, and 101 eq (S +N)/ha/yrto achieve ANC = 0, 5,
10, and 20 (j,eq/L, respectively. The median CL for granitic watersheds based on a critical
ANC limit of 10 j^icq/L was 149 eq/ha/yr.
Target and Dynamic Critical Load Modeling
To simulate the dynamic aspects of damage and recovery, a dynamic model based on
hydrogeochemical processes is required. Dynamic models can be used to simulate soil or
water chemistry or biological response or to calculate a TL within a specified
management timeframe, such as for example the Year 2100. Alternatively, a dynamic
model can be used to calculate a comparable long-term steady-state CL by applying the
model to a point in the distant future. Since the 2008 ISA, dynamic modeling of CLs has
been focused on the Adirondacks, Appalachians, and the Rocky Mountains/Sierra
Nevada (Table 8-8).
8-73
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
Iii the Adirondacks region, the MAGIC model was used by Sullivan et al. (2012a) to
estimate the TLs that would protect the acid-base chemistry of lakes at different ANC
levels. The 117 TLs were calculated for two time periods (2050 and 2100) and three
levels of protection (ANC = 0, 20, and 50 ucq/L) based on the MAGIC model. Results of
the 117 simulated TLs, and associated exceedances, were extrapolated to the regional
population of Adirondack lakes. About 30% of the lakes had TL <500 eq/ha/yrto protect
lake ANC to 50 (.ieq/L. About 600 lakes received S deposition in exceedance of the TL
required to protect to ANC = 50 |ieq/L. in some cases by more than a factor of two.
Based on the model simulations, some critical criteria threshold values were not
obtainable, even when S deposition was decreased to zero (Figure 8-13).
Target Load of S (meqfm ,'y'l

• 425
O 25-50
Ł3 ADK Ecoregiofl
O 50 • 75
j1 State Boundary
O T5-100

• >100

ALSO Sampled Watersheds
ANC Criterion - 50 fjeq/L in 2100
Excectfanec of S Target Load


O No Exeeedanot

P>
ADK Eooregton
O 1-0 to 1.5 !tm«» ft.

a

o 1.51o2.0!im«&fr«
CL
State Boundary
• >20 times the CL



ALSO Sampled Watersheds
ANC Crttenon * SO peq/L m 2100
ADK = Adirondack; ALSC = Adirondack Lakes Survey Corporation; ANC = acid neutralizing capacity; CL = critical load; L = liter;
meq = miiliequivalent; peq = microequivalent.; S = sulfur.
To convert mapped values to units of eq S/ha/yr, multiply by 10.
Source: Sullivan et al. (2012a).
Figure 8-13 Target loads for sulfur deposition in the Adirondack Park to
protect lake acid neutralizing capacity to 50 peq/L in the Year 2010
(left map) and their exceedance (right map).
June 2018
8-74
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Simulations using the dynamic PnET-BGC model (Zhou et al.. 2015c) for the Constable
Pond watershed (a chronically acidified drainage area in the Adirondack Park) suggested
that future decreases in SO42 deposition would be more effective in increasing the lake
water ANC than equivalent decreases in NO3 deposition. Biophysical factors that
affected CLs and TLs included hydrologic residence time. Lakes with longer hydrologic
residence time exhibited less historical acidification and, therefore, could achieve a
higher ANC in response to chemical recovery compared with lakes that had short
hydrologic residence times. This difference was attributed to in-lake production of ANC
in the lakes having long hydrologic residence time. The model outputs further suggested
that forest cutting could enhance acidification by the removal of nutrient base cations
along with the removal of forest biomass. However, forest cutting could also enhance
retention of atmospherically deposited N in an aggrading forest ecosystem.
In another modeling study using PnET-BCG, Zhou et al. (2015b) evaluated lake water
chemistry and the species richness of fish and total zooplankton in 20 Adirondack
watersheds in response to historical acidic deposition and under future deposition
scenarios. Historical acidification in the lakes was related to the total ambient deposition
of S042 + NO3 . the Ca weathering rate, and the simulated preindustrial ANC in the
Year 1850. In the modeled watersheds, changes in the concentration of Al3+ since the
onset of acidic deposition were related to the Ca weathering rate and the hindcast value of
ANC in the Year 1850. Estimated preindustrial ANC for the study lakes ranged from 18
to 190 (ieq/L. Modeling results suggested that lake ANC, fish richness, and total
zooplankton species richness would increase under hypothetical decreases in future acidic
deposition. However, the simulations suggested that biological and chemical recovery
may not be attainable in all of the lakes (Zhou et al.. 2015b).
Total maximum daily load (TMDL) modeling analysis by Fakhraei et al. (2014) for
128 Adirondack lakes designated by New York as acid-impaired under the Clean Water
Act suggested that a further decrease in S deposition of 60% from ambient levels would
allow about 30% of the impaired lakes to achieve the brook trout protection ANC level of
11 (ieq/L by the Year 2050, with an additional 30% recovering by the Year 2200.
Using the same model (PnET-BGC), Zhou et al. (2015a) simulated past and future effects
of N and S acidity on stream chemistry of 12 watersheds in the Great Smoky Mountain
National Park. Three target levels of ANC (0, 20, and 50 j^icq/L) were based on a range of
protection of aquatic life from "minimal" to "considerable." Model simulations suggested
that the level of NO3 + SO42 deposition necessary to achieve a given ANC target was
approximately a linear function of current ANC. Target loads of NO3 + SO42 deposition
for the 12 study streams ranged from 270 to 3,370 eq/ha/yr to reach an ANC of 0 |icq/L
by 2050, 0-2,340 eq/ha/yr to reach ANC of 20 |icq/L by 2050, and 0-1,400 eq/ha/yr to
8-75
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
reach an ANC of 50 |icq/L by 2050. However, the majority of the 12 streams could not
achieve the ANC target of 50 j^icq/L by 2050. This was also true to a lesser extent for the
target of ANC = 20 (ieq/L.
Fakhraei et al. (2016) and Fakhraei et al. (2017a) addressed the likelihood that acidified
streams in Great Smoky Mountains National Park (GRSM) can recover from the adverse
impacts of S and N deposition and restore biotic health, and how long that might take.
They defined the point of harmful effects on the aquatic ecosystems based on converting
pH criteria to ANC then modeling ANC below defined thresholds, using the PnET-BGC
biogeochemical model. The studies updated the previous model application by Zhou etal.
(2015b) by including recent large decreases in S and oxidized N deposition and all
streams in the park listed by Tennessee as water quality impaired under the Clean Water
Act (303(d) listed). Model simulations suggested that TLs varied with measured ANC
(Figure 8-14) and that stream recovery from previous acidification has lagged behind
decreases in S and N deposition due to the dynamics of S adsorption on soils. The TLs
varied with measured ANC in this simulation because the relationship between pH and
ANC is not linear and because stream CO2 can have a strong impact on pH. Simulated
ANC increases were larger per unit decrease in NH4+ deposition than per unit decrease
in S042 or N03 deposition. This finding was attributed to high S adsorption and limited
N retention in watershed soils. Modeling results were extrapolated to other streams
throughout the park. The extrapolation was based on observed linear relationships
between median ANC measured during the period 1993-1996 and the TLs to achieve
various ANC targets. A fixed ANC target of 20 (ieq/L was considered, along with two
other targets that were based on model simulation of preindustrial ANC (ANC in 1850
and ANC 20 j^ieq/L lower than ANC in 1850). The latter target was selected because the
model suggested that ANC in 1850 was generally not obtainable in response to further to
decreases in S and N deposition. Figure 8-15 shows spatial patterns in TL exceedance of
ambient S and N deposition for two endpoint years (2050 and 2150) and two ANC
recovery targets (20 j^ieq/L and 20 j^ieq/L less than modeled preindustrial ANC). For the
303(d)-listed stream watersheds at high elevations within the park critical loads ranged
between 240 and 960 eq/ha/yr of SO42 + NO3 + NH4+ deposition to eight of the twelve
watersheds (Fakhraei et al.. 2017a). For streams in the remaining four watersheds, no
reduction in deposition was sufficient to achieve pH of 6 by 2150 and recovery in these
streams is projected to take centuries.
MAGIC modeling based on simulations of past and future acid-base chemistry of
14 streams in Shenandoah National Park identified a TL of about 188 eq S/ha/yr in the
median modeled stream located on sensitive (siliciclastic) bedrock to achieve
ANC = 50 (ieq/L in 2100. This was 77% lower than the S deposition in 1990 (Sullivan et
8-76
DRAFT: Do Not Cite or Quote

-------
al.. 2008). Many streams had ambient ANC < 20 ueq/L. Hindcast simulations suggested
that preindustrial ANC was above 50 ueq/L in all of the study streams.
+ -r

X

Z
/—v


+
i
1 CO
>>
o

z
1
E
+
"o
1

-------
Target year: 2050
Target ANC: 20 ^molc/L
Target year: 2150
Target ANC: 20 pmoUL
Target year: 2050
Target ANC: Preirdustrial-20 umolc/L
Target year: 2150
Target ANC: Preindustriat-20 pmoyL
o
o
o
0-50
50-200
>200
Source: Fakhraei et al. (2016).
Figure 8-15. Exceedance level of current NOs" + SO42" atmospheric deposition
for 387 stream sites in the GRSM. Exceedances were calculated
for the years 2050 and 2150 using two targets for modeled stream
ANC recovery of 20 |jmolc/L and 20 pmolc/L less than the
simulated preindustrial ANC.
8.5.4.2. International Critical Loads
1	Critical load concepts were initially developed in Europe and only more recently widely
2	applied in the U.S. and Canada. Work on CLs has continued internationally as a basis for
3	setting environmental policy. Some of the recent work conducted outside the U.S. is
4	summarized below. Previous regional CL assessments in Canada have often employed an
5	empirical clay-based soil texture approximation to estimate weathering as input for
6	aquatic and terrestrial steady-state CL modeling Krzvzanowski and Inncs (2010)
June 2018
8-78
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
estimated steady-state CLs, using the SSWC model, and associated exceedances for
protecting freshwater ecosystems in First Nations territory in British Columbia against
water acidification. Estimates of acidic deposition were not in exceedance of the CL at
any of the study lakes.
The 1999 Gothenburg Protocol, under the Long-Range Transboundary Air Pollution
(LRTAP) Convention of the United Nations Economic Commission for Europe
(UNECE) was aimed at reducing the total area of Europe where acidic deposition
exceeded CLs. In 2007, the Coordination Centre for Effects (CCE) of the International
Cooperative Programme for Modeling and Mapping (part of the LRTAP Convention)
issued a call to European countries for results from dynamic models of soil and water
acidification. In response to this call, Norwegian scientists used the MAGIC model to
project recovery of surface waters in Norway. Larssen et al. (2010) described the results
for water quality and fish population status and considered implications for policy
options. The modeling results suggested that surface waters will continue to recover
slowly under existing emissions controls, but about 18% of Norway will still have lakes
that receive acidic deposition that exceeds the CL to protect against aquatic acidification
and in which water quality will continue to be insufficient to support viable populations
of fish and other aquatic organisms.
Both measured and modeled data indicate past acidification and suggest some future
chemical recovery at acid-sensitive locations throughout northern Europe. Some surface
waters have experienced much greater levels of acidification than have acid-sensitive
sites in the eastern U.S., For example, the MAGIC model was used by Skeffington et al.
(2016) to simulate past and future trajectories in stream chemistry of a seriously acidified
small forest watershed in England. Hindcast simulations suggested that ANC decreased
from about 150 to -100 j^ieq/L and pH decreased from 7.1 to 4.2. Hypothesized decreases
in future acidic deposition suggested slow and prolonged chemical recovery over a period
of 250 years to ANC = 43 (ieq/L.
Bishop et al. (2008) applied the SSWC model to lakes in northern Sweden, many of
which are rich in organic matter. They argued that the SSWC model is not suitable for
judging the acidification of individual lakes in regions such as northern Sweden where
the extent of chronic acidification is relatively small and where organic influence on lake
chemistry is pronounced. A variant of the SSWC model predicted preindustrial lake
chemistry at 58 sites where paleolimnological reconstructions of lake chemistry were
available. The authors noted a large discrepancy between SSWC predictions and diatom
reconstructions, and this was attributed largely to short-term fluctuations in modern lake
chemistry.
8-79
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
8.6. Aquatic Acidification Summary and Causal Determinations
In the 2008 ISA, the body of evidence was sufficient to infer a causal relationship
between acidifying deposition and changes in freshwater biota. Overall, the updated
research synthesized in this ISA reflects incremental improvements in scientific
knowledge of aquatic biological effects and indicators of acidification as compared with
knowledge summarized in the 2008 ISA. Studies indicate that aquatic organisms have
been affected by acidification at virtually all trophic levels in sensitive ecosystems and
that these responses have been well characterized for several decades. Research and
observations reported in the 2008 ISA showed consistent and coherent evidence for
effects on aquatic biota, especially algae, benthic invertebrates, and fish that are most
clearly linked to chemical indicators of acidification (pH, ANC, inorganic A1
concentration). Effects on fish species are especially well understood, and many species
have been documented to be adversely affected. Both in situ and lifestage experiments in
fish support previous thresholds of chemical indicators and biological effects.
Physiological perturbations at the organism level of biological organization can lead to
effects on reproduction, growth, and survival. More species are lost with greater
acidification, providing evidence of a biological gradient in effects. Biological shifts such
as loss of acid-sensitive organisms, population declines, and decreased species richness
have been reported from acid-sensitive regions of the U.S. and other countries. Despite
reductions in acidifying deposition, many aquatic ecosystems across the U.S. are still
experiencing effects on ecological structure and functioning at multiple trophic levels.
New information is consistent with the conclusions of the 2008 ISA that the body of
evidence is sufficient to infer a causal relationship between acidifying deposition and
changes in biota, including physiological impairment and alteration of species
richness, community composition, and biodiversity in freshwater ecosystems. Baker
et al. (^Oa) conducted a rigorous review of the effects of acidification on aquatic biota
for the 1990 National Acid Precipitation Assessment Program (NAPAP) State of
Science/Technology reports. In the review, hundreds of laboratory studies, in situ
bioassays, field surveys, whole-system field experiments, and mesocosm studies were
evaluated to provide a synthesis of the effects of acidification on aquatic biota. The
findings in that report, along with literature published after 1990 up to 2007, were
included in the 2008 ISA. The summaries below integrate information known at the time
of the 2008 ISA with newer studies. For most topics related to aquatic acidification, the
fundamental understanding of mechanisms and biological effects has not changed; rather,
additional studies lend further support to the findings from the previous ISA.
The effects of acidification on freshwaters in the U.S. and elsewhere were well
characterized at the time of the 2008 ISA. The sensitivity of a watershed depends on
watershed characteristics such as underlying geology and hydrological flowpaths
June 2018
8-80
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
(Appendix 7). and the sensitivity of species that make up the local biological community.
Changes in biota are linked to chemical indicators in surface water (Appendix 7;
Table 8-9). As stated in the 2008 ISA, biological effects are primarily attributable to low
pH and high inorganic A1 concentration. ANC is also used as a chemical indicator of
acidification because it integrates overall acid-base status and because surface water
acidification models do a better job projecting ANC than pH and inorganic A1
concentrations. However, ANC does not relate directly to the health of biota. The
usefulness of ANC lies mainly in the association between ANC and the surface water
constituents or parameters that directly cause or ameliorate acidity-related stress, in
particular pH, Ca, and inorganic Al.
Acid-sensitive freshwater systems can either be chronically acidified or subject to
periodic episodes of decreased pH and ANC and increased inorganic Al concentration.
Chronically acidic lakes and streams were traditionally defined as having ANC <0 (j,eq/L.
The ANC in freshwater systems where episodic acidification occurs may fall below
0 |icq/L for only a few hours to weeks in a given year (Driscoll et al.. 2001b). During
these acid episodes, water chemistry may exceed acid tolerance of resident aquatic biota.
Biological effects of episodes may include fish mortality, changes in species
composition, and declines in aquatic species richness across multiple taxa. Biological
effects of chronic and episodic acidification have been most clearly documented for
phytoplankton, zooplankton, aquatic invertebrates, and fish and can be linked to changes
in the chemical indicators of aquatic acidification (Table 8-9). An ANC of >50 ueq/L was
used as an indicator for acidification to an ANC level that may harm biota in the U.S.
EPA National Lake Surveys (U.S. EPA. 2009b).
8-81
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Table 8-9 Ecological indicators for aquatic acidification.
Key Indicators
Ecological Effect
Key References
Acid neutralizing
capacity
Commonly set at ANC values less than 0, 20, 50,
and 100 peq/Lto correspond with decreasing levels
of concern (Fiqure 8-4)
Driscoll et al. (2001b), MacAvov and
Bulqer (1995), Baker et al. (1990a),
U.S. EPA (2009b)
Base cation surplus
0 peq/L—risk of inorganic Al leaching to streams
Section 3.2.1.5 2008 Integrated
Science Assessment for Oxides of
Nitrogen and Sulfur-Ecological Criteria
U.S. EPA (2008a) Lawrence et al.
(2007)
PH
<6.0—reduced number of fish species
Driscoll et al. (2001b). MacAvov and
Bulqer (1995), Baker et al. (1990a)
Inorganic Al
>2 pmol/L (54 |jg/L)—toxic to brook trout and likely
other aquatic biota
Baldiqo et al. (2007), Driscoll et al.
(2001b). Wiqinqton et al. (1996a).
MacAvov and Bulqer (1995)
Al = aluminum; L = liter; |jeq = microequivalent; |jg = microgram.
Source: modified from Fenn et al. (2011b).
8.6.1. Phytoplankton
Phytoplankton, photosynthesizing forms of plankton, play important roles in freshwater
systems as primary producers at the base of the aquatic ecosystem food web. These
organisms, encompassing diatoms, cyanobacteria, dinoflagellates, and other algae, vary
in tolerance of acidic conditions. Studies reviewed in the 2008 ISA reported reduced
species richness of phytoplankton with the decreases in pH and increases in inorganic Al
concentrations that are associated with acid-affected surface waters (U.S. EPA. 2008a).
Effects were most prevalent in the 5 to 6 pH range (Baker etal.. 1990a). Since the 2008
ISA, several paleolimnological and field studies have further linked phytoplankton
community shifts to chemical indicators of acidification. For example, Lacoul et al.
(2011) reviewed information on the effects of acidification on plankton in Atlantic
Canada and observed that the greatest changes in phytoplankton species richness
occurred over a pH range of 4.7 to 5.6, just beyond the interval (pH 5.5 to 6.5) where
bicarbonate becomes depleted in the water.
June 2018
8-82
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
8.6.2.
Zooplankton
Zooplankton, the animal forms of plankton, comprise many groups of small freshwater
organisms including protozoans, rotifers, cladocerans, and copepods. Decreases in ANC
and pH and increases in inorganic A1 concentration were shown to contribute to the loss
of zooplankton species and/or abundance in lakes in studies reviewed in the 2008 ISA
(Keller and Gunn. 1995; Schindler et al.. 1985). Possible mechanisms for zooplankton
sensitivity to low pH and ANC include ion regulation failure, reduced oxygen uptake,
inability to reproduce, and inorganic Al toxicity (U.S. EPA. 2008a'). Reported pH
thresholds for zooplankton community alteration generally ranged from 5 to 6 in studies
reviewed in the 2008 ISA. For example, a decrease in pH from 6 to 5 caused decreased
species richness in zooplankton communities in lakes (Holt et al.. 2003; Holt and Yan.
2003; Locke and Sprules. 1994). Newer studies support effects in a similar pH range.
Vinebrooke etal. (2009) reported variations in phytoplankton and zooplankton
communities during a whole-lake experimental acidification of Lake 302S in the
Experimental Lakes area in Ontario. There was a negative effect on zooplankton species
richness as pH decreased from 6.8 to 4.5. Acidification often reduces Ca availability in
lake water and can affect growth and survival of Daphnia spp., an important prey item in
many freshwater food webs (Jeziorski et al.. 2012b). At ANC <0 (ieq/L, zooplankton
richness was low in Adirondack lakes [15 species in highly acidic lakes compared to
35 species at the highest values of ANC in the study (near 200 (j,eq/L) (Sullivan etal..
2006a)].
8.6.3. Benthic Invertebrates
Sediment-associated invertebrates such as bivalves, worms, gastropods, and insect larvae
can be impacted by acidification because H+ and Al can be directly toxic, causing
disruption of ion regulation and reduced reproductive success. As reviewed in the 2008
ISA, decreases in ANC and pH and increases in inorganic Al concentration have been
shown to contribute to the decline in abundance or loss of benthic invertebrates species in
streams. Typically, pH below 5 virtually eliminate all mayflies, a common taxa used to
assess water quality, along with other aquatic organisms from some streams (U.S. EPA.
2008a; Baker and Christensen. 1991). Since the 2008 ISA, a survey of benthic
macroinvertebrates by Baldigo et al. (2009) in 36 streams in the southwestern
Adirondack Mountains indicated that 44 to 56% of macroinvertebrate communities were
severely impacted by acidification at pH <5.1, moderately impacted at pH 5.1 to 5.7, and
unaffected at pH above 6.4 (Baldigo et al.. 2009). Thresholds of pH 5.2 to 6.1 were
identified for sensitive invertebrates from Atlantic Canada; below these pH values,
June 2018
8-83
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
changes in the abundance or presence of invertebrate taxa were observed (Lacoul et al..
2011).
8.6.4. Fish
The primary mechanism that controls the toxic effects of acidification on fish involves
disruption of normal ion regulation at the gill surface, resulting in increased rates of ion
loss, inhibition of ion uptake, and loss of gill function (Bergman et al.. 1988; Wood and
McDonald. 1987; Leivestad. 1982; McWilliams and Potts. 1978). Responses to
acidifying conditions include respiratory and circulatory failure. The effects of low pH,
low ANC, and high inorganic Al concentrations have been well characterized in fish for
many decades. Responses among fish species and lifestages within species to pH and Al
in surface waters have been variable. In general, early lifestages are more sensitive to
acidic conditions than the young-of-the-year, yearlings, and adults (Baker etal.. 1990a;
Johnson et al.. 1987; Baker and Schofield. 1985). Some of the most commonly studied
species include brown trout, brook trout, and Atlantic salmon.
Further characterization of physiological responses (ion regulation, stress responses, gill
Al accumulation) to acidification in fish, mostly Atlantic salmon, trout, and other
salmonids, adds to the existing information on sublethal effects on individual fish species.
Many of the newer studies were conducted in situ and reported varying sensitivity of
different lifestages. Findings are consistent with physiological alterations in fish reported
in the 2008 ISA. Several studies have assessed physiological changes associated with
migratory activities. For example, a recent study assessed gill Al and NKA activity in
smolts moving downstream in well-buffered and acid-impacted migration corridors in the
northeastern U.S. (Kelly et al.. 2015). In the acid-impacted river basin, the fish had
elevated gill Al and lower gill NKA activity.
As summarized in Baker etal. (1990a) and studies reviewed in the 2008 ISA, fish
populations in acidified streams and lakes of both Europe and North America have
declined, and some have been eliminated as a result of atmospheric deposition of acids
and resulting changes in pH, ANC, and inorganic Al concentrations in surface waters.
There is often a positive relationship between pH and number of fish species, at least for
pH values between about 5.0 and 6.5 (Cosby et al.. 2006; Sullivan et al.. 2006a; Driscoll
et al.. 2003b; Bulger et al.. 1999). Additional pH thresholds published since the 2008 ISA
generally reinforce these ranges, and several new studies consider the role of DOC in
modulating pH and subsequent effects on biota. New studies on fish responses to
chemical alarm cues show behavioral effects at pH <6.6. Characterization of ANC and its
levels of concern have not changed appreciably with the newly available information.
June 2018
8-84
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Few or no fish species are found in lakes and streams that have very low ANC (near zero)
and low pH [near 5.0; (U.S. EPA. 2008a; Sullivan et al.. 2006aYI. The number of fish
species generally increases at higher ANC and pH. The pH largely controls the
bioavailability of Al (Driscoll et al.. 2001^. Al is very toxic to fish, and thresholds of
response to elevated concentrations of this metal in acidified waters are summarized in
Table 8-4.
Some of the most in-depth studies of the effects of acid stress on fish have been
conducted in streams in Shenandoah National Park, VA (Cosby et al.. 2006). and lakes in
the Adirondack Mountains, NY (Sullivan. 2015). Effects on fish have also been
documented in acid-sensitive streams of the Catskill Mountains of southeastern New
York and the Appalachian Mountains from Pennsylvania to Tennessee and South
Carolina (Bulger et al.. 2000; Bulger etal.. 1999; SAMAB. 1996; Charles and Christie.
1991V
8.6.5. Thresholds of Response
As reviewed above, new thresholds have been identified in aquatic organisms
(Table 8-10). However, this new information does not appreciably change the
understanding of biological effects associated with chemical indicators or the levels at
which effects occur. Evidence continues to strengthen findings in the 2008 ISA that high
levels of acidification (especially to pH values below 5) eliminate sensitive species from
freshwater streams.
June 2018
8-85
DRAFT: Do Not Cite or Quote

-------
Table 8-10 Results of recent biological effects studies in surface waters
indicative of thresholds of biological response to changes in water
acidity.
Life Forms and Effects
Region
Potential Tipping
Points
Reference
Sensitive invertebrates present
Atlantic Canada
pH 5.2 to 6.1
Lacoul et al. (2011)
Littoral macroinvertebrates
Northern Europe
pH 5.8 to 6.5
Schartau et al. (2008)
Brook trout loss of whole-body Na
Great Smoky Mountains NP
pH 5.1
Neffet al. (2008)
Brook trout loss of whole-body Na
of 10 to 20%
Great Smoky Mountains NP
pH 4.9 to 5.1
Neffet al. (2009)
Fish damage in lakes
Norway
ANC 67 peq/L
Hesthaaen et al. (2008)
Juvenile brown trout mortality in
high DOC streams
Sweden
pH 4.8 to 5.4
Serrano et al. (2008)
Embryo and yolk sac fry survival
during episodes in DOC-rich lakes
Sweden
pH 4.0
Serrano et al. (2008)
Toxicity to brown trout in humic
streams
Northern Europe
pH 5.0; inorganic
aluminum 20 |jg/L
Andren and Rvdin (2012)
Presence of macrophytes
Atlantic Canada
pH 5.0
Lacoul et al. (2011)
Aquatic bird breeding
Atlantic Canada
pH 5.5
Lacoul et al. (2011)
ANC = acid neutralizing capacity; DOC
NP = National Park.
= dissolved organic carbon; L = liter; |jeq = microequivalent; |jg = microgram; Na = sodium;
8.6.6. Biological Recovery
1	Biological recovery (Chapter 1.11.1) can occur only if chemical recovery (Appendix 7) is
2	sufficient to allow growth, survival, and reproduction of acid-sensitive plants and animals
3	(Driscoll et al.. 2001b). Modeling studies in the northeastern and southeastern U.S.
4	suggest that full chemical recovery may take many decades or not occur at all due to the
5	dynamics of S adsorption and desorption and to long-term Ca depletion of soils. As
6	reported in the 2008 ISA, biological recovery lags behind chemical recovery in many
7	aquatic systems, and the time required for biological recovery after chemical recovery is
8	complete is uncertain (U.S. EPA. 2008a). Ecosystems deemed to be on a recovery
9	trajectory are those found to be moving towards a mix of species presence and abundance
10	that approximates the undisturbed state.
June 2018
8-86
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Since the publication of the 2008 ISA, additional studies have assessed recovery of
benthic organisms, although most of this research has been conducted in Canadian and
European waters. New studies continue to support these earlier observations. In general,
recovery of plankton and benthic invertebrates is observed prior to recovery of fish
populations, although most biological communities studied to date have not returned to
preacidification conditions, even after recovery of chemical parameters. In a study
reviewed in the 2008 ISA, zooplankton recovery in experimentally acidified Little Rock
Lake in Wisconsin took one decade, with approximately 40% of the zooplankton species
experiencing a lag time of 1 to 6 years (Frost et al.. 2006). In an experimentally acidified
lake in Canada, zooplankton species diversity partially rebounded to preacidification
levels as pH returned back to 6.1, with rotifers recovering less than crustaceans (Mallcv
and Chang. 1994). Newer studies from the Sudbury Lakes region of Canada indicate
substantial recovery of copepods and cladocerans at pH 5.5 and higher in previously
acidified lakes (Valois etal.. 2011).
In the 2008 ISA, recovery of fish populations following liming or reduction of deposition
was reported in several studies. Newer studies have documented successful
reintroduction of brook trout in previously acidified Adirondack water bodies (Brooktrout
Lake) or recolonization (Honnedaga Lake) by populations in tributary refuges
(Sutherland et al.. 2015; Josephson et al.. 2014). Fish community shifts from historical
acidification have been observed in the upper mainstem of Hubbard Brook (Warren et al..
2008) and other locations.
8.6.7. Most Sensitive and Most Affected Regions
The extent and distribution of sensitive ecosystems and regions in the U.S. were well
known at the time of the 2008 ISA. In the U.S., surface waters that are most sensitive to
acidification based on ANC are largely found in the Northeast, southern Appalachian
Mountains, FL, the upper Midwest, and the mountainous West (McDonnell et al.. 2014b;
Greaver et al.. 2012; Campbell et al.. 2004a; Driscoll et al.. 2001b; Baker etal.. 1990b;
Omernik and Powers. 1983). Levels of acidifying deposition in the West are low in most
areas, acidic surface waters rare, and the extent of chronic surface water acidification that
has occurred to date has been limited (Charles and Christie. 1991). Episodic acidification
does occur in both the East and West at some acid-sensitive locations, and this is part
natural and part human-caused. Geographic patterns in acidification sensitivity vary in
response to spatial differences in geology, hydrologic flow paths, presence and depth of
glacial till, climate, and others. Sullivan (2017) mapped the locations of known lakes and
streams that had low ANC across the country (Figure 8-11).
June 2018
8-87
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
8.6.8.
Critical Loads
Since the 2008 ISA, considerable CL research has been conducted in the U.S. New
generalized empirical CL estimates include 571 eq N/ha/yr in the Northeast and
286 eq N/ha/yr in the West for episodic acidification of high elevation lakes under
high-flow conditions (Baron et al.. 201 lb). Heard et al. (2014) estimated CL = 74
eq/ha/yr to protect against chronic acidification in high-elevation lakes in the Sierra
Nevada. Steady-state CLs have been derived at many locations since the 2008 ISA.
Steady-state CLs of S and N for lakes in the Adirondack Mountains (1,620 eq/ha/yr) and
for the central Appalachian streams (3,700 eq/ha/yr) were calculated for maintaining a
surface water ANC of 50 |icq/L on an annual basis (NAPAP. 2011). Sullivan et al.
(2012b) calculated CL values in the Blue Ridge ecoregion for maintaining stream ANC at
50 (ieq/L. The calculated CLs were less than 500 eq/ha/yr at one-third of the study sites.
Observations showed that about one-half or more of the stream length in the study region
was in exceedance of the CL of S deposition for protecting aquatic resources to an
ANC = 50 (ieq/L over the long term. McDonnell et al. (2014b) calculated steady-state
aquatic CLs to protect southern Appalachian Mountain streams against acidification to
ANC = 50 (ieq/L and other critical benchmark values. Results showed that nearly
one-third of the stream length in the study region (mainly streams in Virginia, West
Virginia, North Carolina, and Tennessee) had a CL of S deposition <500 eq/ha/yr, which
was less than the estimated regional average S deposition (600 eq/ha/yr). Critical loads
for acid deposition to lakes in Class I and II wilderness areas of the Sierra Nevada were
estimated in 2008 to protect ANC to 0, 5, 10, and 20 (j,eq/L levels, which span the range
of minimum ANC values observed in those lakes. Median CLs were 217 (ANC = 0), 186
(ANC = 5), 157 (ANC = 10), and 101 (ANC = 20) eq/ha/yr. The median CL for granitic
watersheds based on a critical ANC limit of 10 j^icq/L was 149 eq/ha/yr. It was estimated
that slightly more than one-third of the lakes received acidic deposition higher than their
CL.
In addition to the steady-state and empirical CLs described above, CL estimates are
available from dynamic modeling. NOs leaching in stream water in California was both
simulated (by the DayCent model) and determined empirically to be approximately
1,214 eq N/ha/yr (Fenn et al.. 2008). Zhou et al. (2015a) simulated past and future effects
of N and S on stream chemistry of 12 watersheds in the Great Smoky Mountain National
Park. Three target levels of ANC (0, 20, and 50 j^icq/L) were used, based on a range of
protection of aquatic life from minimal to considerable. TLs of NO3 + SO42 deposition
for the 12 study streams ranged from 270 to 3,370 eq/ha/yr to reach ANC = 0 |icq/L by
2050, 0 to 2,340 eq/ha/yr to reach ANC = 20 (j,eq/L by 2050, and 0 to 1,400 eq/ha/yr to
reach ANC = 50 |icq/L by 2050. However, the majority of streams could not achieve the
ANC target of 50 j^icq/L. This was also true to a lesser extent for the target of
June 2018
8-88
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
ANC = 20 (ieq/L. Modeling studies also suggested that complete recovery may not be
possible in the Appalachian Mountains (Sullivan et al.. 201 lb). For some sites, one or
more of the selected critical ANC levels (0, 20, 50, 100 j^icq/L) could not be achieved by
2100, even if S deposition was decreased to zero and maintained at that level throughout
the simulation. MAGIC modeling based on simulations of past and future acid-base
chemistry of 14 streams in Shenandoah National Park identified a target load of about
188 eq kg S/ha/yr to achieve ANC = 50 (ieq/L in 2100 in the median modeled stream
located on sensitive (siliciclastic) bedrock, which was 77% lower than the S deposition in
1990 (Sullivan et al.. 2008). Many streams had ambient ANC <20 (ieq/L, although
hindcast simulations suggested that preindustrial ANC was above 50 j^ieq/L in all of the
study streams.
In the Adirondack Mountains, TLs were calculated for two time periods (2050 and 2100)
and three levels of protection (ANC = 0, 20, and 50 j^ieq/L). Results of simulated TLs,
and associated exceedances, were extrapolated to the regional population of lakes. About
30% of the lakes had TL <500 eq/ha/yr to protect lake ANC to 50 j^icq/L (Sullivan et al..
2012a). Also in the Adirondack Mountains, Zhou et al. (2015c) ran simulations using the
PnET-BGC model, which suggested that future decreases in SOr deposition would be
more effective in that region in increasing the lake water ANC than equivalent decreases
in N03 deposition. In another modeling study of 20 Adirondack watersheds, lake ANC
and fish and total zooplankton species richness were projected to increase under
hypothetical decreases in future acidic deposition, but model projections suggested that
lake ecosystems may not achieve complete chemical and biological recovery in the future
(Zhou et al.. 2015b). Estimates of preindustrial ANC for the study lakes ranged from 18
to 190 (ieq/L. The magnitude of simulated historical acidification represented by ANC
loss ranged from about 26 to 100 (ieq/L.
8-89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
APPENDIX 9. BIOLOGICAL EFFECTS OF
FRESHWATER NITROGEN
ENRICHMENT
This appendix characterizes the biological effects of nitrogen (N) nutrient enrichment
from atmospheric deposition to freshwater systems. Biogeochemical processes and
chemical indicators associated with nutrient enrichment of fresh waters are discussed in
Appendix 7. Atmospheric deposition constitutes only a portion of total N load in many
water bodies; however, it may be the dominant source of N in some remote aquatic
ecosystems, such as headwater and lower order streams and alpine lakes, which are more
affected by N deposition than other sources of N. Appendix 9J_ presents an overview of
freshwater nutrient enrichment in these systems and includes a discussion on the
characteristics of water bodies sensitive to N deposition, the effects of N deposition on
nutrient limitation, phosphorus (P) interactions, and climate modification of N response.
Inputs of P have direct implications for how ecosystems respond to N deposition, and
recent trends in increased atmospheric deposition of P are discussed in this context.
Indicators of biological responses to nutrient enrichment (Appendix 9.2) effects of N on
species diversity, ecosystem structure, and function (Appendix 9.3) and emerging
research on the links between nutrient enrichment and animal behavior and disease
(Appendix 9.4) provide a basis for identifying thresholds of biological response
(Appendix 9.5) and determining causation based on new information and evidence from
prior N assessments (Appendix 9.6).
9.1. Introduction to Nitrogen Enrichment and Eutrophication in
Freshwater Systems
In the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur—Ecological Criteria (2008 ISA), the body of evidence was sufficient to infer a
causal relationship between N deposition and the alteration of species richness, species
composition, and biodiversity in freshwater ecosystems (U.S. EPA. 2008a). Increased
atmospheric N inputs to freshwater systems via runoff or direct deposition, especially to
N limited and N and phosphorus (P) colimited systems, can stimulate primary
productivity (Figure 9-1). Eutrophication is the process of enriching a water body with
nutrients resulting in increased growth and change in the composition of primary
producers (algae and/or aquatic plants). One of the consequences of eutrophication is low
oxygen levels in the water body when these primary producers decompose. Changes in
biological indicators of N enrichment, including chlorophyll a, phytoplankton
June 2018
9-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
(free-floating algae) biomass, periphvton (algae attached to a substrate) biomass, and
diatoms (major algal group with cell walls made of silica), provide evidence for N
effects. The transport of atmospheric N inputs downstream can exacerbate eutrophic
conditions in higher order streams, rivers, and lakes. Atmospheric deposition to
watersheds affects processes along the freshwater-to-ocean continuum, including coastal
and estuarine systems (Appendix 10). New information is consistent with the conclusions
of the 2008 ISA that the body of evidence is sufficient to infer a causal relationship
between N deposition and changes in biota, including altered growth and
productivity, species richness, community composition, and biodiversity due to N
enrichment in freshwater ecosystems.
Source: Modified from Baron et al. (2011 b).
Figure 9-1 Conceptual model of the influence of atmospheric nitrogen
deposition on freshwater nutrient enrichment.
June 2018
9-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
Based on the studies described throughout Appendix 9 and in the 2008 ISA, the
freshwater ecosystems in the U.S. most likely to be sensitive to N deposition are
headwater streams, lower order streams, and alpine lakes, which have very low nutrients
and productivity and are far from local pollution sources (U.S. EPA. 2008a'). Even small
inputs of N in these water bodies can increase nutrient availability or alter the balance of
N and P, which can stimulate growth of primary producers and lead to changes in species
richness, community composition, and diversity. Remote mountain lakes in the western
U.S. are naturally oligotrophic and are considered among the aquatic ecosystems most
sensitive to N deposition (Williams et al.. 2017b). A portion of these lakes and streams in
the western U.S. are in Class I wilderness areas (Williams et al.. 2017b; Clow et al..
2015; Nanus et al.. 2012). Survey data and fertilization experiments from studies
reviewed in the 2008 ISA have documented increases in algal productivity as well as
species changes and reductions in diversity at high-elevation lakes in the western U.S. in
response to increased availability of N (U.S. EPA. 2008a). Some examples include the
Snowy Range in Wyoming, the Sierra Nevada, Lake Tahoe, and the Colorado Front
Range. In studies reviewed in the 2008 ISA, atmospheric N deposition (approximately 1
to 5 kg N/ha/yr) can cause an increase in phytoplankton and periphyton biomass in some
alpine lakes. For example, in both the Beartooth Mountains of Wyoming and the Rocky
Mountains of Colorado, N deposition as low as 1.5 kg N/ha/yr affects algal productivity
(Baron. 2006; Saros et al.. 2003). The responses of high-elevation lakes can vary
considerably depending on catchment characteristics and the amount of deposition
(Appendix 9.1.1).
With increased characterization of nutrient inputs in remote high-elevation lakes and
streams, fate and transport processes, N and P dynamics, phytoplankton response, and
downstream effects in coastal/estuarine systems, the understanding of the role of N in
freshwater eutrophication has evolved in recent decades. As conveyed in the 2008 ISA,
the historical emphasis on P as a major cause of freshwater eutrophication was based on a
number of highly influential studies in the 1960s and 1970s that focused on the role of
wastewater, in particular phosphate detergents, in causing excessive algal blooms in Lake
Mendota (WI), Lake Washington (WA), Lake Erie, and many other locations
(Edmondson. 1991; Schindler. 1974; Schindler et al.. 1971; Edmondson. 1969;
Vollenweider. 1968; Hasler. 1947). These observations guided approaches to freshwater
lake and stream ecology for many years. Over time, an increasing recognition that N
inputs can stimulate phytoplankton growth under certain conditions, coupled with further
understandingof aquatic biogeochemistry (Appendix 7) and of the connectivity between
freshwater and receiving estuaries and coastal waters, has led to recommendations to
consider both N and P in nutrient reduction strategies (Dodds and Smith. 2016; Gobler et
al.. 2016; Paerl et al.. 2016b; Lewis etal.. 2011; Scott and McCarthy. 2010; Conlev et al..
2009; Paerl. 2009; Lewis and Wurtsbaugh. 2008).
9-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
There is increasing understanding of the frequency and shifts in the types of nutrient
limitation in various systems. In oligotrophic clear water lakes where atmospheric N
inputs are low and light is not limiting, phytoplankton primary production is typically N
limited (Bcrgstrom et al.. 2015; Hcsscn. 2013). In the 2008 ISA, results from surveys,
paleolimnological reconstructions, experiments, and meta-analyses of hundreds of studies
have shown N limitation to be common in fresh waters, especially in remote areas, and a
nearly universal eutrophication response to N enrichment in lakes and streams that are N
limited (U.S. EPA. 2008a; Elser et al.. 2007; Bergstrom et al.. 2005; Elseretal.. 1990).
Tropical and subtropical lakes, and lakes having small watersheds relative to the lake
surface/volume, also tend to be N limited (Paerl and Scott. 2010; Elser et al.. 2007). As
reported in the 2008 ISA, some alpine lakes have exhibited shifts from N limitation to
between N and P limitation or to P limitation (U.S. EPA. 2008a'). In a meta-analysis
reviewed in the 2008 ISA, Elser et al. (2007) found that N limitation occurred as
frequently as P limitation in freshwater ecosystems. Newer studies published since the
2008 ISA add to the evidence that reservoirs, rivers, and freshwater lakes can exhibit N
limitation and N and P colimitation (Paerl et al.. 2014; Lewis etal.. 2011; Conlev et al..
2009; Sterner. 2008). N limitation appears to be increasingly common in freshwater
systems, probably because their nutrient dynamics are being altered significantly by
growing agricultural and urban P inputs (Paerl et al.. 2016b; Grantz et al.. 2014; Paerl et
al.. 2014; Finlav et al.. 2013).
A source of P to remote water bodies is deposition of dust. In an analysis of data from the
U.S. EPA National Lakes surveys and National Rivers and Streams surveys Stoddard et
al. (2016) found continent-wide increases in TP between 2000 and 2014, especially at
sites that exhibited low disturbance (Appendix 9.1.1.2). Among the large natural P
emission sources are soils, vegetation, and biomass combustion ash. Other notable
sources include industry, agriculture, and mining. Although P is not a criteria pollutant,
inputs of P may contribute to eutrophication and effect shifts in lake trophic status from P
to N limitation or to colimitation. In addition, Stoddard et al. (2016) observed that TN
was strongly correlated with TP in lakes and streams on a national scale. Although the
authors determined that TP was increasing at "minimally disturbed sites," they observed
that TN was not increasing at those sites. Responses of aquatic ecosystems to
atmospheric N deposition are heavily dependent on surface water P concentrations. Thus,
because P inputs can alter N response, the impact of recent trends in increased P
deposition is important to consider when evaluating nutrient status in water bodies
sensitive to atmospheric inputs (Appendix 9.1.1.4).
9-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
9.1.1.
Deposition to Freshwater Systems
Both N and P inputs to fresh waters sensitive to atmospheric deposition can affect
nutrient limitation. The composition of N deposition is shifting in the U.S. from oxidized
to reduced forms of N, with implications for the receiving systems. Atmospheric
deposition of P may affect lake response to N inputs, and recent trends point to
widespread increases in P deposition in the U.S. (Stoddard et al.. 2016) and globally
(Brahncv et al.. 2015; Tipping et al.. 2014) (Appendix 9.1.1.2).
9.1.1.1. Nitrogen Deposition Sources and Trends
Sources of N and trends in atmospheric deposition are described in Appendix 2. Briefly,
N is deposited in various reduced and oxidized forms, including organic N. Deposition
can be wet (rain or snow), or dry. Atmospheric deposition of reduced N has increased
relative to oxidized N in parts of the U.S. including the East and Midwest in the last few
decades, shifting from a NOs dominated to a NH4 dominated condition, and this trend is
expected to continue under existing emissions controls (Li et al.. 2016d; Pinderetal..
2008; U.S. EPA. 2008a). In the soil or water, much of the deposited NFU+ is either taken
up by biota or nitrified to NO3 and leaches to water bodies mainly as NO3 (Hcsscn.
2013). Up to 70% of deposited NO3 in remote alpine lakes in the western U.S. is
anthropogenic in origin, with the largest sources being atmospherically delivered
fertilizers from agriculture (approx. 60%) and fossil fuel combustion (approx. 10%)
(Hundev et al.. 2016). N deposition to snow and glaciers are important sources of N to
alpine lakes and streams that are fed by meltwaters. For example, approximately 50% of
lakes in Glacier National Park and 10-20% of lakes in the central Rockies receive glacial
meltwater (Saros et al.. 2010). At high-elevation sites like those in the Rockies and Sierra
Nevadas, N deposition estimates are uncertain, especially for dry deposition
(Appendix 9.5).
Moving from lower order streams to higher order streams, atmospheric N from direct
deposition, runoff, and leaching from terrestrial ecosystems combines with other diffuse
and point sources of N. The contribution from other terrestrial sources ofN, such as
fertilizer, livestock waste, septic effluent, and wastewater treatment plant outflow, often
becomes much more important in downstream than in upland areas. About 75% of N
inputs are retained in the watershed or denitrified, and 25% are exported to surface waters
regardless of the dominant N input, including deposition (Howarth et al.. 2012; Howarth
ctal.. 1996a). Table 7-1 summarizes studies quantifying N deposition contribution to
total N loading in U.S. freshwater systems.
June 2018
9-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
9.1.1.2.
Characteristics of Freshwater Systems Sensitive to Atmospheric
Deposition of Nitrogen
Various factors affect the sensitivity of remote water bodies to atmospheric deposition.
These factors include the spatial and temporal patterns of nutrient limitation and the
physical and chemical attributes of the catchment (Williams et al.. 2017b; Hundev et al..
2014; Nanus et al.. 2012; U.S. EPA. 2008a'). Thus, the same amount of N deposition can
lead to different N loading depending on the characteristics of the receiving watershed
and water body (Hcsscn. 2013; Bergstrom. 2010).
In high-elevation lakes above the tree line in areas with steep slopes, sparse vegetation,
exposed bedrock, and shallow rocky soils, changes in productivity and biodiversity of
algal assemblages can occur with little or no lag time (Baron et al.. 201 lb). The
hydrology of these systems is dominated by spring snowmelt (Spaulding et al.. 2015).
Seasonal meltwaters can deliver a nutrient pulse to alpine lakes and streams, with glacial
meltwater contributing more NOs than snowmelt does (Slemmons et al.. 2015;
Slemmons et al.. 2013; Saros et al.. 2010; Baron et al.. 2009). Other freshwater systems,
such as some nonalpine freshwater lakes, reservoirs, and rivers that exhibit N limitation
and N and P colimitation, are sensitive to additional N inputs (Paerl et al.. 2016b; Paerl et
al.. 2014; Lewis et al.. 2011; Conlev et al.. 2009; Elser et al.. 2009b; Sterner. 2008). N
from atmospheric deposition represents a proportion of total N in these systems, although
agricultural and wastewater inputs are often predominant.
In the 2008 ISA, lakes and streams with high concentrations of NO, . indicative of
ecosystems being N saturated (Appendix 7.1.2.1). were reported at a variety of locations
throughout the U.S., including the San Bernardino and San Gabriel mountains within the
Los Angeles Air Basin (Fcnn and Poth. 1999; Fenn et al.. 1996; Riggan et al.. 1985). the
Front Range of Colorado (Williams et al.. 1996a; Baron et al.. 1994). the Allegheny
Mountains of West Virginia (Gilliam et al.. 1996). the Catskill Mountains of New York
(Stoddard. 1994; Murdoch and Stoddard. 1992). the Adirondack Mountains of New York
(Wigington et al.. 1996b). and the Great Smoky Mountains in Tennessee (Cook et al..
1994). All of these regions, except the Colorado Front Range, received more than about
10 kg N/ha/yr atmospheric deposition ofN throughout the 1980s and 1990s. The Front
Range of Colorado received up to about 5 kg N/ha/yr of total (wet + dry) deposition
(Sullivan et al.. 2005). less than half of the total N deposition received at many of these
other locations. Since the 2008 ISA, several long-term monitoring studies have shown
temporal decreases in surface water NO3 concentration corresponding to decreases in
atmospheric N deposition (Appendix 7.1.2.1). These regions include the Appalachian
Mountains, the Adirondacks, and the Rocky Mountains (Driscoll et al.. 2016; Kline et al..
2016; Strock et al.. 2014; Eshleman et al.. 2013; Elser et al.. 2009b; Bergstrom and
June 2018	9-6	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Jansson. 2006). NOs in extremely high concentrations can have direct adverse effects on
fish, as well as invertebrates and amphibians. These effects are observed at
concentrations much higher than would commonly be attributable to atmospheric
deposition, are not included in this ISA, and were not defined as a primary biological
indicator in the 2008 ISA.
9.1.1.3. Nitrogen Deposition Effects on Nutrient Limitation
The 2008 ISA reported on gradient studies of undisturbed northern temperate, mountain,
or boreal lakes that receive low levels of atmospheric N deposition. These studies found
strong relationships between N limitation and primary productivity where N deposition
was low and between P and N + P limitations where N deposition was high (Bcrgstrom
and Jansson. 2006; Bcrgstrom et al.. 2005; Fenn et al.. 2003a'). As reviewed in the
2008 ISA, a comprehensive survey of 42 unproductive lakes (oligotrophic lakes with TP
less than or equal to 25 j^ig/L) along gradients of N deposition in Europe and North
America showed increased inorganic N concentration and productivity to be correlated
with atmospheric N deposition (Bcrgstrom and Jansson. 2006). Unproductive lakes
receiving low atmospheric N deposition were N limited (<2.5 kg N/ha/yr). At about 2.5
to 5 kg N/ha/yr, colimitation of N and P was observed, while at N deposition above
5 kg N/ha/yr, the lakes were P limited. Based on the study findings and paleolimnological
evidence, Bcrgstrom and Jansson (2006) suggested that most lakes in the northern
hemisphere may have originally been N limited and that atmospheric N deposition has
changed the balance of N and P in lakes.
Studies published since the 2008 ISA have continued to characterize nutrient
relationships and evaluate the potential for N deposition to contribute to the
eutrophication of water bodies (Table 9-1). Consistent with the 2008 ISA findings,
research literature after 2007 indicates that N deposition is correlated with a shift from N
to P limitation in certain high-elevation water bodies (Hessen. 2013). Elser et al. (2009b)
conducted a nutrient-limitation study across a gradient of lakes in the Rocky Mountains
of Colorado that receive low (<2 kg N/ha/yr; N = 20) and high (>6 kg N/ha/yr; N = 16) N
deposition. Nutrient enrichment bioassays indicated that P limitation or colimitation was
prevalent in the population of high-deposition lakes (9 of 16 lakes), while only one of the
lakes showed N limitation. In contrast, only 4 of 20 low-deposition lakes showed P
limitation. Based on relative response ratios (RR; chlorophyll concentration in a given
treatment normalized to chlorophyll in control), chlorophyll responded more strongly to
N relative to P in low-deposition lakes where N limitation was stronger (Figure 9-2).
These data were included in Elser et al. (2009a) who reported that lakes in Norway,
Sweden, and Colorado affected by N deposition showed a similar pattern. In lakes with
June 2018
9-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
high N deposition, phytoplankton was predominately P limited, whereas in lakes with
low N deposition, N limitation was more common.
Based on an analysis of lake water chemistry data from 106 alpine lakes in Sweden,
Slovakia, Poland, and the Rocky Mountains of Colorado and nutrient bioassays from high
mountain lakes in Sweden and the Rocky Mountains of Colorado, Bergstrom (2010)
documented how N:P mass ratios (TN:TP and DIN:TP) vary in oligotrophic lakes located
in northern Europe and the Rocky Mountains. Nutrients (N, P, N + P, control) were
added to water from 28 unproductive lakes. The majority of the lakes were N limited and
the shift from N to P limitation was strongly affected by N deposition. More than half
(54%) of the oligotrophic study lakes had a TN:TP mass ratio <25. A DIN:TP ratio of 1.5
was indicative of an N limited lake while a ratio of 3.4 was P limited. The DIN:TP ratio
was a better indicator than the TN:TP ratio for nutrient limitation of phytoplankton
because TN may include a large fraction of biologically unavailable N.
9-8
DRAFT: Do Not Cite or Quote

-------
Table 9-1 Summary of studies using diatoms as biological indicators of nitrogen enrichment evaluated in the
literature since the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria.
Study Site
Ambient N Deposition
Diatom Response
Species
Reference
Three national parks 0.25 ±0.15 to
in Washington State
(Mt. Rainier, North
Cascades, and
Olympic)
Overall, only one lake (Hoh Lake) of 10 study lakes where cores A. formosa, F.
1.10 ±0.21 kg/ha/yr NH4+-N;
0.34 ± 0.04 to
1.32 ±0.19 kg/ha/yr NO3--N;
0.59 ± 0.07 to
2.42 ± 0.27 kg/ha/yr inorganic
N
were collected showed clear evidence of impacts from N
deposition based on changes in sediment diatom communities.
crotonensis, and
F. tenera
Sheiblev et al.
(2014)
Western U.S.
(Glacial National
Park, Greater
Yellowstone
Ecosystem, and
eastern Sierra
Nevada)
Sierra Nevada = 2 kg/ha/yr
(current) to 4 kg/ha/yr (early
1990s) total wet inorganic N
deposition.
GNP = 0.5-1.5 kg/ha/yr
(1980-2006).
GYE = 0.5-1.1 kg/ha/yr
(1980-2006)
A critical load of 1.4 kg N/ha/yr wet deposition changed diatom
community structure in both the eastern Sierra Nevada and the
Greater Yellowstone Ecosystem, although N deposition rates
between the two regions and the timing of diatom community
shifts were different. No diatom community changes were
observed in Glacier National Park lakes.
A. formosa and
F. crotonensis
Saros et al. (2011)
Rocky Mountains
Central Rockies NO3 and
NhU+ = 1.4-2.5 kg Nr/ha/yr. N.
Rockies = 2.0-3.4 kg Nr/ha/yr.
Lakes fed by snowpack meltwater had greater sediment diatom
taxonomic richness over the last century (35 to 54 taxa)
compared with lakes that were fed by both glacial and snowpack
meltwater (12 to 26 taxa) which are higher in NO3".
(-50% of Glacier National Park lakes and -0-20% of lakes in
central Rockies receive glacial meltwater).
Multiple
Saros et al. (2010)
Rocky Mountains
Wet deposition ranged from
0.8 to 3.2 kg N/ha/yr at study
sites.
High abundance of key indicator species of N enrichment	Multiple
(A. formosa, F. crotonensis) in lakes with low to moderate NO3"
precluded an attempt to quantify the surface water NO3"
concentrations that elicit diatom community changes in
high-elevation lakes using a diatom-based transfer function.
Arnett et al. (2012)
June 2018
9-9
DRAFT: Do Not Cite or Quote

-------
Table 9-1 (Continued): Summary of studies using diatoms as biological indicators of nitrogen enrichment
evaluated in the literature since the 2008 Integrated Science Assessment for Oxides of
Nitrogen and Sulfur-Ecological Criteria.
Study Site	Ambient N Deposition	Diatom Response	Species	Reference
Rocky Mountains >3 kg N/ha/yr	A maximum growth rate of 0.2 ± 0.04/day at 0.5 |jM N for the A. formosa	Nanus et al. (2012)
diatom A. formosa was measured following N addition to identify
a NO3" threshold for this species in surface water.
Central Rocky
Mountains/Beartooth
Mountains located
along the Montana
and Wyoming
borders
Not specified
Fossil-diatom assemblages from two lakes near each other (one
snowpack-fed and one glacier-fed) indicated greater diatom
assemblage turnover in the glacier-fed lake. The glacier-fed lake
showed evidence for mild N enrichment starting approximately
1,000 yr ago, and the increase in abundances of A. formosa and
F. crotonensis occurred much earlier, suggesting glacial
meltwater was a source of N.
A. formosa and F.
crotonensis
Slemmons et al.
(2015)
Central Rocky
Mountains/Beartooth
Mountains located
along the Montana
and Wyoming
borders
Not specified
Diatom species richness from the top of sediment cores was
1.8x higher in snowpack-fed lakes compared with glacier- and
snowpack-fed lakes, whereas richness did not differ between
core bottoms (ca. 1850) of the lakes. Compared with snow-fed
lakes, N enriched glacial snowpack-fed lakes were dominated by
A. formosa and F. crotonensis.
A. formosa and
F. crotonensis
Slemmons and
Saros (2012)
Grand Teton
National Park
Estimated at 2.5 kg N/ha/yr
Directional change in benthic diatom assemblages after 1960
that is correlated with atmospheric deposition.
Benthic diatoms
Spauldina et al.
(2015)
Uinta Mountains, UT 0.02-0.04 kg km2
Four of five lakes with recent increase in productivity (between
1940 and 1960) show comparable shifts in diatom community
composition linked to atmospheric deposition of N and P. Lake
sediment records show an increasing abundance of nitrophilous
A. formosa.
A. formosa
Hundev et al.
(2014)
Mount Rainier, North 0.38-3.24 kg N/ha/yr (wet)
Cascades, and
Olympic national
parks, WA
20 taxa of phytoplankton responded to N enrichment. F. tenera
and F. crotonensis are recommended as indicators of N
enrichment in the Pacific NW. The threshold for phytoplankton
biomass growth was 13-25 |jg DIN/L.
Multiple
Williams et al.
(2016a)
Northern Cascade
Mountains, WA
1.1 to 3.4 kg N/ha/yr
No significant difference in diatom biovolume or phytoplankton
community structure between snow-fed lakes and glacial
snowpack-fed lakes.
Multiple
Williams et al.
(2016b)
June 2018
9-10
DRAFT: Do Not Cite or Quote

-------
Table 9-1 (Continued): Summary of studies using diatoms as biological indicators of nitrogen enrichment
evaluated in the literature since the 2008 Integrated Science Assessment for Oxides of
Nitrogen and Sulfur-Ecological Criteria.
Study Site
Ambient N Deposition
Diatom Response
Species
Reference
National parks of the
western Great Lakes
(Superior and
Michigan) region
1.5-5 kg N/ha/yr (1980-2010) In 63% of study lakes, change in diatom community correlates
with lake sediment 815N, which in turn relates to measured Nr
inputs (deposition and indirect watershed inputs). In about 36%
of the lakes, sediment 815N was statistically correlated to some
form of deposited Nr. A number of such lakes are in watersheds
believed to have high N retention associated with shallow
subsurface flow during snowmelt. The authors consider this to be
the primary path for NO3" transport. Diatom community change
is also expected to be related to nutrient inputs indirectly by
runoff and climate variability.
Multiple, including
F. crotonensis
and A. formosa
Hobbs et al. (2016)
U.S.
Not specified
Shifts in diatom community composition away from N intolerant Multiple
species.
Pardo et al. (2011a)
North America and
Greenland
0.8 to 2 (<50°N) to <0.5
(>50°N)
A meta-analysis of 52 alpine, Arctic, and boreal montane lakes
showed increased beta diversity in the 20th century. Observed
diatom assemblage turnover in alpine and Arctic lakes is related
to temperature changes, while in mid-latitude alpine lakes
changes are linked to N deposition.
Multiple
Hobbs et al. (2010)
Baptiste Lake,
Alberta, Canada
Not specified
Lake sediment cores indicate prevalence of diatom assemblages Multiple
that favor nutrient-rich conditions for at least the last 150 yr.
From -1980 to present, a distinct increase in Stephanodiscus
hantzschii was observed.
Adams et al. (2014)
George Lake,
Killarney Provincial
Park, Ontario,
Canada
6-7 kg N/ha/yr (2013)
Relative abundance of A. formosa increased over a 20-yr period
while lake total N concentration and regional N deposition
decreased. Increases in A. formosa were directionally related to
increases in air and epilimnetic temperatures.
Multiple, including
A. formosa
Sivaraiah et al.
(2016)
Whitefish Bay,
Ontario, Canada,
also meta-analysis
of 200 lakes in North
America and Europe
<5 to <1 kg N/ha/yr
No effects of N deposition were observed on diatom community Multiple	Ruhland et al.
structure in Whitefish Bay. Observed shifts in Arctic lakes (1850) planktonic and (2008)
and temperate lakes (1970) were in response to climate change, benthic diatoms
Ha = hectare; kg = kilogram; km = kilometer; |jM = micromole; N = nitrogen; NH4+-N = nitrogen as ammonium; N03 = nitrate; P = phosphorus; yr = year.
June 2018
9-11
DRAFT: Do Not Cite or Quote

-------
IX
cc
cc
cc
Low
High	Low
Atmospheric N deposition
High
N = nitrogen; P = phosphorus.
The response ratio (RR) has no units. (A) Response to P enrichment alone (RR-P); (B) response to N enrichment alone (RR-N);
(C) response to combined N and P enrichment (RR-NP); (D) relative response to N vs. P (RR-N/RR-P, equivalent to final chl N/final
chl P). For RR-N/RR-P, a value >1 indicates stronger N limitation while a value <1 indicates stronger P limitation. Error bars
indicate ± SE. The result of t tests comparing low- and high-deposition means for each parameter are given in each panel.
Source: Elser et al. (2009^
Figure 9-2 Phytoplankton responses to nitrogen and/or phosphorus
enrichment for Rocky Mountain lakes receiving low or high
atmospheric nitrogen deposition, given as the ratio of final
chlorophyll concentration in the enriched treatment
(+phosphorus, +nitrogen, or +nitrogen + phosphorus) to the
chlorophyll concentration in the unenriched control.
1	N deposition gradient studies conducted in Sweden support findings of shifts from N to P
2	limitation in lakes. Liess et al. (2009) compared a population of lakes in northern Sweden
3	with lower N deposition (N = 7) to lakes in the south with higher deposition (N = 6).
4	Atmospheric N deposition showed a strong positive correlation with TN. In the sampled
5	lakes, northern lakes received 2 to 6 kg N/ha/yr, while southern lakes received 10 to
6	12 kg N/ha/yr. Sampling of epilithic communities indicated a weakly positive correlation
7	of atmospheric N deposition to epilithon N:P ratios, suggesting that epilithic communities
June 2018
9-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
were generally more N limited in the northern lakes and more P limited in the southern
lakes, which received higher N deposition. Bergstrom et al. (2008) evaluated
phytoplankton responses to N and P enrichment in unproductive Swedish lakes along a
gradient of atmospheric N deposition. They found that regional and seasonal patterns in
nutrient limitation were related to the level of N deposition. In the south, high N
deposition was accompanied by high lake DIN concentrations during early summer, with
subsequent P limitation. Later in the summer, DIN concentrations declined, and lakes
switched to N and P colimitation, and then to N limitation. Bergstrom et al. (2008)
concluded that N limitation is probably the natural state of these unproductive lakes and
that P limitation has been induced by increased N availability caused by atmospheric N
deposition.
Other studies in high alpine systems have shown that N deposition does not necessarily
have a consistent effect on N or P limitation. In an assessment of 29 alpine lakes and
ponds in the Canadian Rocky Mountains (Banff National Park, Alberta and Yoho
National Park, British Columbia), certain ponds appeared to be N limited while most
water bodies sampled did not show evidence of N limitation (Murphy et al.. 2010).
DIN: TP ratio analysis indicated N limitation in only 14% of the sampled water bodies. N
limitation was observed more frequently in shallow ponds than lakes, and during the late
summer sampling period, the ponds had a significantly lower mean DIN:TP ratio. In
another study from the Rocky Mountains, surface sediment samples and surface water
N03 concentrations were collected from an N deposition gradient (1 to 3.2 kg Nr/ha/yr
in wet deposition) across 46 high-elevation lakes to characterize diatom community
changes (Arnett et al.. 2012). The researchers found that even in lakes with NOs
concentrations below quantification (<1 (ig/L). diatom assemblages were already
dominated by key indicator species characteristic of moderate N enrichment, and the
authors were unable to identify a threshold. They noted that small inputs of reactive N
can shift diatom species composition along the length of the NO3 gradient and that the
lakes switch from N limited oligotrophy to P limitation. In a study of 11 lakes in northern
subarctic Sweden situated along an altitudinal/climate gradient with low N deposition
(<1 kg N/ha/yr), the lakes did not show a shift from mainly N to mainly P limitation
reported from U.S. lakes where N deposition was higher. However, mainly P limitation
was observed in high alpine lakes with rocky catchments and high DIN:TP ratios, while
mainly N and N + P colimitation was occurring in subalpine and lower mid-alpine lakes
(Bergstrom et al.. 2013). The authors attributed these observations to climate and
catchment characteristics and suggested warming would have a greater impact than N
deposition on the subset of sparsely vegetated high alpine lakes, in contrast to the
majority of subarctic lakes with N or N + P colimitation, where both warming and N
inputs will alter phytoplankton response.
9-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
9.1.1.4. Phosphorus Deposition Interactions
P deposition has direct implications for how ecosystems respond to N deposition.
Although this appendix is focused on the biological effects of N enrichment, inputs of P
may effect shifts in lake trophic status from P to N limitation or to colimitation and
increase the total nutrient supply to water bodies affecting how the system responds to N.
Since the 2008 ISA, several meta-analyses have reported an increase in P deposition to
water bodies (Stoddard et al.. 2016; Brahnev et al.. 2015; Tipping et al.. 2014). This
recent evidence suggests that the predominantly dry deposition of fine (<10 p.) and coarse
(<100 |a) particulates ("dust") containing P play a role in the enrichment effects of N
deposition to fresh waters and their catchments. Data from the U.S. EPA National Lakes
Assessment and National Rivers and Streams Assessment were analyzed by Stoddard et
al. (2016) to determine whether total P (TP) concentrations changed between 2000 and
2014. The authors found continent-wide increases in TP, especially at sites with low
disturbance. Median stream TP concentrations increased from 26 |ig/L (2000-2004) to
56 |ig/L (2013-2014); median lake TP increased from 20 |ig/L (2007) to 37 |ig/L (2012).
From the 2004-2014 surveys, the percentage of stream length where TP was <10 |ig/L
decreased from 24.5 to 1.6; the percentage of lakes where TP was <10 |ig/L decreased
from 24.9 to 6.7 between 2007 and 2012. Additional research has corroborated the
findings of Stoddard et al. (2016) in other locations (Zhu et al.. 2016a; Brahnev et al..
2015; Tipping et al.. 2014) and has investigated shifts from N to P limitation or
colimitation, as well as P deposition's role in prolonging N limitation (Appendix 9.2.2).
In a regression analysis of deposition and water quality data in 700 upland lakes across
21 alpine regions globally, Brahnev et al. (2015) suggested that P deposition may play a
large role in alpine lake trophic status. The authors evaluated the strength in N, P, and
N:P relationships in deposition and in lakes. Their 5-year, Community Atmospheric
Model (CAM4) simulation modeling indicates that P deposition may have increased
globally by 1.4 times the preindustrial deposition rate. TIN:TP of deposition and in lakes
showed a strong relationship (r2 = 0.82,p< 0.0001) with P deposition and lake water
(r2 = 0.64,p < 0.0001) and N deposition and lake water (r2 = 0.20,p < 0.05). Where the
deposition's molar ratio of N:P was less than 20, intermittent or persistent N limitation
was observed to be induced atmospherically. The authors concluded that the chemistry of
atmospheric deposition influences nutrient limitation of alpine oligotrophic lakes by
changing nutrient ratios as well as increasing the absolute supply of nutrients to these
ecosystems. A global-scale analysis by Tipping et al. (2014) supported the observation by
Brahnev et al. (2015) about oligotrophic lakes and indicated the importance of accounting
for sustained P deposition to assess the impact of productivity of anthropogenically
emitted N deposition. The authors analyzed data on P deposition measured globally (82%
of sites were in Europe and North America) from 1954-2012 at 250 sites. They found
June 2018
9-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
that "oligotrophic lakes, tropical forests, and ombrotrophic peatlands" are the most likely
to experience significant effects of atmospheric P deposition increase.
Zhu et al. (2016a) examined P deposition in China's forest, grassland, desert, lake, marsh,
and karst ecosystems and confirmed that China's wet P deposition in terrestrial
ecosystems (0.21 kg P ha/yr) is comparable to Tipping et al. (2014) TP deposition
estimates (2013 ambient mean wet deposition = 13.69 ± 8.69 kg N/ha/yr with fluxes
ranging from 0.47 to 47.71 kg N/ha/yr). The average N:P ratio for wet deposition in
China was reported as 77:1. The mean annual N concentration at 41 monitoring sites
correlated linearly with precipitation (r2 = 0.120; p = 0.0027), and N deposition also
correlated significantly with rainfall (r2 = 0.217./; = 0.002). The authors suggested that
the high N:P ratios in atmospheric wet deposition could shift systems to P limitation,
influencing ecosystem structure and function.
9.1.1.5. Climate Modification of Ecosystem Response to Nitrogen
Nutrient inputs to fresh waters are occurring within the context of physical, chemical, and
biological modifications caused by increased annual mean temperature and magnitude of
precipitation associated with climate change (Greaver et al.. 2016). Projected shifts in
runoff and timing and quantity of flushing will alter local and regional hydrology
(Whitehead et al.. 2009). Nutrient loads to surface waters are expected to increase due to
predicted increases in surface water flow (Adrian et al.. 2009; Whitehead et al.. 2009).
Air temperature increases will lead to warmer surface waters, altering the thermal
stratification of water bodies and affecting the community composition and distribution
of aquatic biota in streams and lakes (Adrian et al.. 2009; Keller. 2007). The increased
surface water temperatures will increase the rate of algal growth (Whitehead et al.. 2009).
In regions with no evidence of increased atmospheric nutrient inputs, warming trends are
observed to enhance competitiveness of planktonic diatoms like Asterionella formosa,
which are typically associated with elevated N, indicating that climate change has
significant direct and indirect effects on algal species composition. Climate change may
thus enhance the effects observed in areas with nutrient increases alone (Ri'ihland et al..
2015). Appendix 13 includes a more detailed discussion of how climate (e.g., temperature
and precipitation) modifies ecosystem response to N loading.
9.2. Biological Indicators
Increased NOs in surface water is a chemical indicator of freshwater N nutrient
enrichment (Appendix 7.1.2.1). Many of the critical loads discussed in Appendix 95_ are
June 2018
9-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
based on NOs in surface water. A critical load is a quantitative estimate of exposure to
one or more pollutants below which significant harmful effects on specified sensitive
elements of the environment do not occur according to present knowledge KSprangcr et
al.. 2004; Nilsson and Grennfelt. 1988); Chapter 1.2.2.31. Biological indicators of
freshwater N enrichment discussed in the 2008 ISA included chlorophyll a,
phytoplankton and periphyton biomass, and changes in lake nutrient status (U.S. EPA.
2008a). Paleolimnological records of shifts in diatom community composition were also
used to assess the effects of N deposition. Dose-response relationships between N and
biological indicators were reported in the 2008 ISA, and the new literature continues to
support these findings. Diatom community shifts (Appendix 9.2.1). and phytoplankton
biomass nutrient limitation shifts (Appendix 9.2.3) have been used as a basis for
determining critical loads for nutrient enrichment (Appendix 9.5). These same biological
indicators are discussed further below along with new studies. In the current review, the
response of microbial enzymes to N and P and increased incidence of harmful algal
blooms (HABs) in freshwater habitats are discussed as possible biological indicators of
the effects of elevated N.
9.2.1. Diatoms
Diatoms are commonly used to monitor environmental conditions in water bodies over
time. While many diatom studies explain patterns of variation in community composition
in relation to environmental (Sniol and Stoermer. 2010). as well as spatial factors
(Soininen et al.. 2016; Vilmi et al.. 2016). few studies have used experimental methods to
address the processes underlying the patterns (Smol and Stoermer. 2010; Pither and
Aarssen. 2006). Changes in diatom species assemblages in paleolimnological studies of
mountain lakes disturbed only by atmospheric deposition and climate change were
reported in the 2008 ISA. Chlorophytes, such as A. formosa and Fragilaria crotonensis,
generally prefer high concentrations of N and are able to rapidly dominate the flora when
N concentrations increase (U.S. EPA. 2008a; Findlav et al.. 1999). These two
nitrophilous species of diatom are used as biological indicators of the effect of N in water
bodies; however, increased relative abundance of A. formosa has also been attributed to
lake warming in some regions where N deposition is decreasing (Sivaraiah et al.. 2016).
Table 9-1 summarizes diatom studies published since the 2008 ISA. The majority of
these studies highlight the observed influence of 20th century N deposition on diatoms,
causing an increasing abundance of nitrophilous diatoms such as A. formosa and F.
crotonensis. Appendix 93_ considers the effects of N deposition on diatom diversity.
Thresholds of diatom response are reported in Appendix 9.5.
June 2018
9-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
9.2.2.
Ratios of Nitrogen and Phosphorus
Trophic status is a way to characterize bodies of water in terms of their productivity.
Increasing deposition of N to water bodies shifts element ratios, which in turn affect algal
growth, diversity, and community structure (Appendix 9.3.2). Algal species have
different nutrient optima for growth and cellular uptake of N, and P alters the elemental
composition of primary producers at the base of the freshwater food web. When N is the
limiting nutrient, atmospheric deposition of N will increase benthic algal (Liess et al..
2009) and phytoplankton (Hessen. 2013; Elser et al.. 2010) N:P ratios. Nutritional
responses of aquatic ecosystems to atmospheric N deposition are heavily dependent on
surface water P concentrations, which may also be affected by atmospheric P inputs
(Appendix 9.1.1.4). Thus, various chemical ratios of N to P can be useful for evaluating
biological responses of water bodies affected by deposition (Table 9-1).
In the 2008 ISA, several studies reported N:P ratios in which a shift in nutrient limitation
was observed. When DIN:TP values are greater than reference values, growth
stimulation, N and P colimitation, or P limitation commonly occur (Sickman et al.. 2003).
In a Swedish lake survey reviewed in the 2008 ISA, N limitation occurred in lakes where
the DIN:TP mass ratio was less than 7 (DIN concentrations <33 (j,M). Colimitation of N
and P was found in lakes with DIN:TP ratio between about 8 and 10, and P limitation
occurred at DIN:TP values greater than 10 (Bcrgstrom et al.. 2005). Other thresholds for
N limitation were reported in the literature to occur at DIN:TP ratios <4 (Lohman and
Priscu. 1992) and <10 (Wold and Hershev. 1999). Bergstrom et al. (2005) reported an
index (DIN: [chlorophyll a:TP]) to assess the eutrophication of lakes in response to N
deposition.
Hundev et al. (2014) assessed trophic status for six remote alpine lakes in the Uinta
Mountains, UT based on TP, TN, chlorophyll a, Secchi depth, and N:P relationships. One
of six lakes was P limited, two were N limited, one varied by month, and the limitation
was uncertain in the other two. Trophic status was difficult to determine in some of the
lakes because the ratios were values that could indicate either N or P limitation depending
on the threshold used, and some lakes were on the boundary between oligotrophic and
mesotrophic. In Green Lake in Colorado's Front Range, the ratio of DIN:TP in the
epilimnion over the course of the study was consistently above 4 and averaged
16.3 ± 2.76, suggesting the lake was P limited (Gardner et al.. 2008). Nutrient ratios
(TN:TP, DIN:TP, and NOs :TP) were compared to nutrient bioassays to evaluate the use
of these ratios in predicting the limiting nutrient. There was agreement in 29% of the
bioassay results, suggesting that nutrient ratios are not the best predictor of nutrient
limitation in this subarctic region.
June 2018
9-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
In an in situ mesocosm bioassay study of two subalpine lakes in the Sierra Nevada, CA,
Heard and Sickman (2016) modeled the effective dose of N addition along an N gradient
(N as KNO3 and NaNCh ranging from 0 to 50 (imol/L) with P addition as KH2PO4 held
constant at 1.5 (imol/L. DIN:TP ratios in the two lakes were <4.0. Given that DIN:TP
ratios <0.6 indicate N limitation and DIN:TP ratios > 0.6 indicate intermediate limitation,
both lakes were deemed N limited. The authors observed that the P addition delayed the
shift to P limitation (prolonged N limitation) for phytoplankton. Effective doses for N for
the two lakes were lower than the other study lakes receiving N addition only. They
acknowledged the spatial and temporal variation in lake P concentrations, as well as
concern for increased P deposition.
In a comparison of diatom-inferred lake nutrient records and surface water monitoring
data (collected since the early 1980s) from Baptiste Lake in the Canadian Rockies,
Adams et al. (2014) found total Kjehldahl nitrogen (TKN) was a significant predictor of
chlorophyll a, where chlorophyll a was independent of TP. They observed that the lake
experienced eutrophic conditions for >150 years based on diatom sediment core data.
Diatom-inferred TKN data tracked current TKN dynamics measured in the water column.
9.2.3. Phytoplankton Biomass Nitrogen (N) to Phosphorus (P) Limitation Shift
As described in Appendix 9.1.1.3. lakes may shift from N limitation to P limitation with
elevated N deposition and this can be assessed considering the RR-N compared to the
RR-P I Figure 9-2; (Elser et al.. 2009b)l. Phytoplankton biomass growth may shift from N
to P enrichment, altering lake primary production (Hcsscn. 2013). Williams et al. (2017a)
used the RR-N/RR-P to define a biological threshold of RR-N/RR-P = 1 above which
phytoplankton biomass P limitation is more likely than N limitation and developed
critical loads for western U.S. mountain lakes (Appendix 9.5).
9.2.4. Periphyton/Microbial Biomass
Periphyton mats are biofilms of algae, cyanobacteria, fungi, microinvertebrates, organic
detritus, inorganic particles, and heterotrophic microbes imbedded within a matrix and
attached to submerged substrates in aquatic systems (e.g., stream or lake bottoms). In the
2008 ISA, no studies reported resource requirements for periphyton, although several
papers described stimulated growth with N amendments in ecosystems throughout the
U.S. (Annex C of the 2008 ISA), including streams in Alaska, Arizona, Iowa, Texas,
Minnesota, and Missouri and lakes in California, Colorado, and Massachusetts (U.S.
EPA. 2008a). Additional lake bioassay experiments that enriched the water column down
June 2018
9-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
into the sediments found enhancement of periphyton growth on bioassay container walls
in experiments in California, Wyoming, and Massachusetts (Smith and Lee. 2006;
Nvdick et al.. 2004; Axler and Reuter. 1996). Strong N limitation of benthic algae has
also been inferred in streams of Arizona (Grimm and Fisher. 1986). California (Hill and
Knight. 1988). Missouri (Lohman et al.. 1991). and Montana (Lohman and Priscu. 1992).
Since the 2008 ISA, few studies on N effects on periphyton biomass have been identified.
In Ditch Creek, WY, a relatively undisturbed stream with N accumulation primarily from
high N2 fixation, biofilm growth on rocks increased after snowmelt and N2 fixers
dominated the algal assemblage (Kunza and Hall. 2014). A shift to non-N2 fixing taxa
was observed later in the season. In this stream with little to no anthropogenic influence,
N fixation was higher than denitrification. Burrows et al. (2015) assessed microbial
respiration and biomass using nutrient-diffusing substrata in 20 boreal streams in
Sweden. An increase in N availability led to increased microbial activity. Microbial
biomass was primarily N limited, and distinct microbial communities were associated
with inorganic N in stream water. In a series of nutrient-diffusing substrata studies across
rivers in the U.S. mountainous West, arid West, and the Midwest, Reisingeretal. (2016)
assessed nutrient limitation patterns of benthic biofilms. Regarding regional differences,
there was little evidence of N limitation in midwestern rivers due to high nutrient
concentrations, while nutrient limitation of biofilms was common in the summer in
mountainous western rivers. Increasing developed lands decreased the probability of
nutrient-limited river biofilms. In a meta-analysis of nutrient-diffusing substrate studies
from North America Beck et al. (2017). broad spatial factors such as ecoregion described
most of the variation in nutrient limitation. Variables affecting algal biomass response to
N included land use, riparian canopy cover, the presence of soluble reactive P, and
season.
Microbial communities involved in plant litter decomposition in streams have been
shown to be altered by nutrient concentrations. Most studies have examined the effects of
N and P in combination. However, Fernandes et al. (2014) conducted a series of N
addition studies in microcosms to assess leaf litter decomposition with increasing N
concentration and temperature in streams. In general, increased temperature led to an
increase in decomposer activity and a decrease in the amount of N needed. Field studies
in six low-order streams spanning an N gradient in the Ave River basin, Portugal,
suggested that eutrophication modulates leaf litter decomposition processes (Lima-
Fernandes et al.. 2015). Leaf litter diversity synergistically affected leaf litter
decomposition while biomass fungal and invertebrate decomposers increased with stream
eutrophication status but decreased in the most eutrophic of the six streams. Dunck et al.
(2015) reported decreased primary production and leaf litter decomposition in highly
eutrophic streams and streams with little human influence compared to streams that were
9-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
intermediate along the trophic gradient. Kominoski et al. (20151 observed that microbial
litter breakdown rates increased across low to moderate nutrient enrichment gradients in
experimental stream channels where N:P ratios were varied to allow for examination of N
effects. A meta-analysis of plant matter decomposition in streams by Ferreira et al.
(2015) suggested that the effects of nutrient inputs might be strongest in oligotrophic
streams due to the low background nutrient concentrations and high magnitude of
nutrient enrichment in these systems.
9.2.5. Chlorophyll a
The concentration of chlorophyll a, a pigment present in all photosynthetic organisms, is
a common measure of algal productivity and is hence an easily documented biological
indicator of change in aquatic ecosystem productivity. Quantification of chlorophyll a is
one approach for estimating phytoplankton biomass. Chlorophyll a is used as the primary
indicator of trophic state in the U.S. EPA National Lakes Assessment and as a water
quality indicator in many state and federal monitoring programs (U.S. EPA. 2016h.
2009b). Chlorophyll a is being used as an indicator of nutrient enrichment in U.S. EPA's
National Nutrient Program lYU.S. EPA. 1998b); Appendix 7.1.61. The U.S. EPA is
working with the states to develop numeric nutrient criteria to better define levels of N
and P that affect U.S. waters. The numeric values include both causative (N and P) and
response (chlorophyll a, turbidity) variables to assess eutrophic conditions. The U.S.
EPA's National Nutrient Program to reduce eutrophication in water bodies developed
recommended nutrient criteria for rivers and streams in 14 U.S. ecoregions for the states
to use as a starting point to develop their own criteria (U.S. EPA. 1998b).
Surveys and fertilization experiments reported in the 2008 ISA show increased inorganic
N concentration and aquatic ecosystem productivity, as quantified by chlorophyll a
concentration, to be strongly related (U.S. EPA. 2008a). At the time of the 2008 ISA,
increases in lake phytoplankton biomass (as chlorophyll a) with increasing N deposition
were reported in several regions, including the Snowy Range in Wyoming (Lafrancois et
al.. 2003b) and across Europe (Bergstrom and Jansson. 2006). A meta-analysis of
enrichment bioassays in 62 freshwater lakes of North America reported in the 2008 ISA
found algal growth enhancement from N amendments to be common in slightly less than
half of the studies (Elseretal.. 1990). There was a mean increase in phytoplankton
biomass of 79% in response to N enrichment (Elseretal.. 1990). This meta-analysis was
repeated, incorporating study sites from multiple countries and a much larger data set,
with similar results (Elser et al.. 2007). Chlorophyll a continues to be a common
biological indicator of N nutrient enrichment in the research literature from 2008 to
present. Dodds and Smith (2016) conducted a review of N and P dynamics in stream
June 2018
9-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
ecosystems and concluded that both N and P are strongly correlated to chlorophyll and
algal biomass. In the western U.S., Elser et al. (200%) examined chlorophyll a response
in Rocky Mountain, CO lakes where atmospheric deposition ranged from 2 to
7 kg N/ha/yr. Concentrations of chlorophyll were 2 to 2.5 times greater in
high-deposition lakes relative to low-deposition lakes.
Since the 2008 ISA, nutrient threshold values for chlorophyll a responses have been
identified for a subset of western mountain lakes. Williams et al. (2016a) calculated an
average DIN threshold of 13 |ag DIN/L for increased chlorophyll a concentration across
nine western mountain lakes (25 |_ig DIN/L for increased chl a concentration beyond
interannual variation). More than 25% of 207 lakes in Mount Rainier and the northern
Cascades exceeded the 13 p.g DIN/L threshold. Heard and Sickman (2016) determined
threshold values for phytoplankton growth for Sierra Nevada lakes. The doses of N that
characterize the initial phytoplankton growth response (10% effective dose), rapid
phytoplankton growth (50% effective dose), and saturating nutrient level (90% effective
dose) in early- and late-season conditions within two high-elevation lakes were
established. These doses were then compared with monitoring data from lakes within
Yosemite, Sequoia, and Kings Canyon national parks. The range of threshold values for
the 10-50% effective doses were in the range of 0.3 to 4 (.iniol N/L (5-56 |_ig N/L). The
50% effective doses were exceeded by 18% (late season) to 29% (early season) of
monitored lakes, suggesting that N inputs via atmospheric deposition affect
phytoplankton in many lakes in this region. The lakes that exceeded effective doses
tended to be at high elevation on steep, north-facing slopes with limited vegetative
growth.
Chlorophyll a can also be used as a bioindicator of lake response over decadal and longer
periods. Hundev et al. (2014) took sediment cores from six remote alpine lakes in Utah to
assess trends in lake productivity. In five of the six lakes, sedimentary chlorophyll a and
percentage of organic matter was observed to be relatively constant from the beginning of
the record (mid 1800s using 210Pb dating; 1200 to 1800 estimated by linear regression)
until 1940-1960 when lake production progressively increased.
As reported in the 2008 ISA and discussed further in Appendix 7.1.2.9. dissolved organic
carbon (DOC) affects acidity and N cycling and is increasing in some U.S. surface waters
(Monteith et al.. 2007; Evans et al.. 2006). Recent studies indicate different
phytoplankton responses to N and dissolved organic matter (DOM) inputs depending on
nutrient status of the lakes and background DOC (Deininger et al.. 2017a; Daggett et al..
2015). Daggett et al. (2015) selected a low DOC, N and P colimited water body (Jordan
Pond in Acadia National Park, ME) and an N limited lake with higher DOC (Sargent
Lake in Isle Royale National Park, MI) to assess the effects of an N gradient on algal
9-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
biomass following addition of DOM. Both of these Class I areas have a similar N
deposition rates (6-7 kg total inorganic N/ha/yr wet deposition). DOM addition
stimulated phytoplankton biomass in both lakes regardless of nutrient limitation or
background DOC concentration. Phytoplankton biomass in N limited Sargent Lake
increased with both N and DOM inputs, while in N and P colimited Jordan Pond, algae
were sensitive primarily to DOM addition. Appendix 9.3.2.3. discusses a whole-lake N
fertilization study along a gradient of DOC.
Not all studies have shown a positive relationship between N inputs and increased
chlorophyll a. In the Canadian Rockies, Murphy et al. (2010) performed nutrient
enrichment bioassays on 29 water bodies in Banff and Yoho national parks using
1 mg N/L additions. They found that N amendment did not significantly increase final
total phytoplankton chlorophyll concentration across all of the water bodies sampled
during either early or late summer 2007. Although phytoplankton collected from ponds
showed several responses to nutrient amendment, the effect of N on total chlorophyll did
not differ significantly between pond and lake communities. Harvested final total
chlorophyll concentrations from the bioassays were significantly higher for ponds than
lakes. In water bodies in Wapusk National Park, Manitoba, a significant linear
relationship occurred between chlorophyll a and TP across all ponds but not between
chlorophyll a and total N [TN; (Svmons et al.. 2012)1. In another study in Colorado's
Front Range in Green Lake Valley (part of the Niwot Ridge Long-Term Ecological
Research Project), Green Lake 4, a well-studied lake in the area, was sampled weekly
throughout the ice-free summer period to assess chlorophyll response (Gardner et al..
2008). Epilimnetic chlorophyll a concentrations peaked at 4 (ig/L in late July. In the same
study, N03 addition to an in situ mesocosm on the lake (930 (ig/L NOs . for a final
exposure of 1,240 (ig/L NO3 given background concentrations) did not increase algal
biomass significantly in comparison with the control, while phytoplankton chlorophyll a
increased in P and P + NO3 additions, indicating the lake was P limited during the
summer.
Although hypothesized to be driven by industrial point source emissions, an increase in
aquatic primary production in some lakes in the Athabasca Oil Sands Region in Alberta,
Canada appears to follow regional patterns of annual and seasonal changes in
temperature. In a comparison of spectrally inferred chlorophyll profiles of lake sediments
in 23 diverse undisturbed lakes, Summers et al. (2016) observed higher chlorophyll a
concentrations in surface sediments compared with concentrations inferred for the period
before oil sands development, irrespective of lake morphology, landscape position, N
deposition, or other limnological characteristics. Spatial and temporal patterns in inferred
chlorophyll a were positively correlated with annual and seasonal changes in temperature
9-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
and were consistent across the region, suggesting a regional impact on productivity rather
than impacts associated with point source emissions.
Some research has been conducted regarding the biological impacts of changes in N
retention. Heini et al. (2014) examined chlorophyll a at various depths in a lake in
southern Finland, and found that algae and cyanobacteria can affect water chemistry
instead of just responding to it. Differences in phytoplankton strongly affected observed
short-term differences in chemical properties. The chemical conditions in the deeper
waters during summer were generally more stable than in the layers near the surface of
the lake.
Since the 2008 ISA lake surveys, fertilization experiments and nutrient bioassays
continue to show a strong correlation between N inputs and chlorophyll a. Determination
of additional thresholds for chlorophyll a in remote high-elevation lakes have further
linked this indicator to atmospheric inputs of N. The relative importance of DOC in
modulating lake response to nutrients is better understood than at the time of the
2008 ISA (Appendix 7.1.2.9). The use of chlorophyll a as an indicator in national lake
and stream assessments and for state numeric nutrient criteria (Appendix 7.1.6) is further
supported by studies reported in this ISA.
9-23
DRAFT: Do Not Cite or Quote

-------
Table 9-2 Summary of additional evidence for nitrogen limitation on productivity of freshwater ecosystems
that has been evaluated since the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria.
Region
N Deposition
N Addition
Biological Effects
Reference
Sierra Nevada,
CA
Not specified
Gradient of N (as
KNOs and
NaNC>3) ranging
from 0 to
50 pmol/L
(dose-finding
study), then range
used to model
biological
response was
only up to about
18 pmol/L
10, 50, and 90% effective doses for excess phytoplankton growth were
calculated for two high-elevation lakes. The modeled doses were compared
with lake chemistry monitoring data to assess nutrient status. The 50%
effective doses were exceeded by 18% (late season) to 29% (early season)
of monitored lakes, suggesting that N inputs via atmospheric deposition
affect phytoplankton in many lakes in this region. The threshold for
stimulation of phytoplankton (10 to 50% ED) was 5-56 |jg N/L.
Heard and Sickman
(2016)
Rocky Mountains, 2-7 kg N/ha
CO
None
Atmospheric N deposition increased the stoichiometric ratio of N and P in
lakes. Phytoplankton growth in 16 high N deposition lakes (~7 kg N/ha) was
P limited whereas in 20 low N deposition lakes (~2 kg N/ha) growth was
primarily N limited.
Elseret al. (2009a)
Rocky Mountains,
CO
High >6 kg N/ha/yr;
low <2 kg N/ha/yr
(NADP)
Enrichment of	Phytoplankton response to increased inputs of N was inferred from
7.5 pmol/L N (as chlorophyll changes in bioassay data from 20 low N deposition lakes and
NH4NO3) for N,	16 high N deposition lakes. Concentrations of chlorophyll and seston C
there was also a	were 2-2.5 times higher in the high N deposition lakes relative to the low N
P and N + P	deposition lakes, while high-deposition lakes also had higher seston C:N
treatment	and C:P (but not N:P) ratios.
Elseret al. (2009b)
Rocky Mountains,
CO
Not specified
Added 930 pg/L
NO3"; with
background,
exposure was
1,240 pg NO3-;
added 93 pg/L
TDP
In in situ mesocosm experiments with water from Green Lake 4, chlorophyll Gardner et al. (2008)
a did not increase significantly with addition of NO3" in comparison with the
control, while P and N + P treatments resulted in increases.
June 2018
9-24
DRAFT: Do Not Cite or Quote

-------
Table 9-2 (Continued): Summary of additional evidence for nitrogen limitation on productivity of freshwater
ecosystems that has been evaluated since the 2008 Integrated Science Assessment for
Oxides of Nitrogen and Sulfur-Ecological Criteria.
Region
N Deposition
N Addition
Biological Effects
Reference
Utah
Not specified
None
Using sediment core data from six remote alpine lakes, chl a and
percentage of organic matter are relatively constant from the beginning of
the record until 1940-1960 in five of six lakes, when production
progressively increased. The authors found the limiting nutrient difficult to
identify.
Hundev et al. (2014)
Alberta, Canada Not specified
None
TKN was a significant predictor of chl a in Baptiste Lake where chl a was
independent of TP measured in the water column.
Adams et al. (2014)
Wapusk National Not specified
Park, Canada
N (NH4NO3) and
P (KH2PO4) were
added to increase
the nutrient
concentrations by
10x mean
ambient
concentrations
Nutrient enrichment bioassays were conducted on water samples from
21 lakes and ponds with chl a concentrations from 1.2 to 10.8 |jg/L. In total,
13% of the lakes and ponds were found to be N limited, 26% P limited, 26%
colimited, and 38% did not respond to either N or P additions.
Svmons et al. (2012)
Sweden
Gradient rates of N
deposition ranging
from 100 to
1,000 kg N/km2/yr
in four regions
during summers of
2004-2006
N: 1 mg/L and/or
the
concentrations of
P by 100 |jg/L
(molar N:P ratio,
23:1)
Phytoplankton were N limited in northern lakes (low N deposition), while in
the southern lakes higher N deposition was accompanied by increased lake
DIN concentrations and a switch from N to P limitation. N limitation was
common during the summer. As summer progressed, P limitation in these
systems switched to duel and colimitation of N and P, and to N limitation,
due to exhaustion of the DIN pool in the lakes.
Berqstrom et al. (2008)
Sweden
<1 kg N/ha/yr
Nutrients were
added to increase
[N] as NH4NO3 by
100 |jg/L
(7.2 pmol/L)
and/or [P] as
KH2PO4 by
10 |jg/L
(0.3 pmol/L)
In phytoplankton nutrient addition bioassays using water from
high-elevation lakes, phytoplankton was subject to P limitation and became
increasingly N and NP colimited at lower elevation. Chlorophyll
concentrations in the bioassays were lower with increasing elevation and
this pattern held over the whole growing season.
Berqstrom et al. (2013)
June 2018
9-25
DRAFT: Do Not Cite or Quote

-------
Table 9-2 (Continued): Summary of additional evidence for nitrogen limitation on productivity of freshwater
ecosystems that has been evaluated since the 2008 Integrated Science Assessment for
Oxides of Nitrogen and Sulfur-Ecological Criteria.
Region
N Deposition N Addition
Biological Effects
Reference
Sweden
2 to 12 kg/ha/yr None
N deposition positively related to total N and total P. The highest proportion
of N fixing cyanobacteria (although only consisting of 5% of the algal
biovolume) was found where N deposition was lowest. Epilithic periphyton
N:P ratios increased with higher N availability from deposition.
Liess et al. (2009)
C = carbon; ED = effective dose; KH2P04 = monopotassium phosphate; N = nitrogen; NADP = National Atmospheric Deposition Program; NH4N03 = ammonium nitrate;
N03" = nitrate; P = phosphorus; TDP = total dissolved phosphorus; TKN = total Kjehldahl nitrogen; TP = total phosphorus.
June 2018
9-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
9.2.6. Potential Biological Indicators
In addition to widely used biological indicators of nutrient enrichment (chlorophyll a,
periphyton/microbial biomass, diatoms), freshwater HABs and enzymes are altered by N
availability. Formation of HABs are more relevant downstream where multiple sources of
N contribute to elevated nutrient levels sufficient for bloom formation. Only a limited
number of studies have used enzymes to date.
9.2.6.1. Harmful Algal Blooms
Freshwater HAB formation has become more prevalent in the U.S. in recent decades
(Lopez et al.. 2008). coinciding with increased N in surface waters. Although risk of
bloom formation remains low for high-elevation oligotrophic water bodies with
atmospheric N as the dominant source, atmospheric N may combine with other sources to
contribute to total N loading in downstream lacustrine and riverine systems. Freshwater
HABs can affect taste and odor of drinking water, lead to hypoxic conditions, impact
recreational uses of surface waters, and produce toxins harmful to humans, domestic
animals, and wildlife (Lopez et al.. 2008). The major harmful algal group in freshwater
environments are the cyanobacteria (blue-green algae). Cyanobacterial toxins are
produced by several genera including Microcystis, Anabaena, Nodularia,
Aphanizomenon, Cylindrospermopsis, and Oscillatoria. These toxins may target the liver
(hepatotoxins such as cylindrospermopsins and microcystins), the nervous system
(neurotoxins such as anatoxins and saxitoxins), or the skin (dermatoxins).
There is evidence of the role for N in freshwater HABs in water bodies across the U.S. In
western Lake Erie, cyanobacterial growth was N limited during bloom conditions in late
summer, and the N fixing cyanobacterium Anabaena became dominant following the
observed N limitation (Chaffin et al.. 2013). In a survey of eutrophic midcontinent lakes,
microcystins (common cyanobacterial toxins associated with HABs), were detected in all
blooms (Graham et al.. 2010). In another U.S. survey, TN was strongly correlated to
microcystin concentration in lakes and reservoirs (Beaver et al.. 2014). The highest
cyanobacteria abundance was observed in the Midwest where agriculture is a dominant
land use, and TN concentrations are higher. Using data from the U.S. EPA National
Lakes Assessment (U.S. EPA. 2009b). Yuan et al. (2014) modeled a threshold for the
probability of occurrence of Microcystis, a common non-N2 fixing cyanobacterial genus.
In their analysis, the frequency of occurrence of high microcystin concentrations
depended most strongly on TN, with weaker associations to chlorophyll a. The calculated
June 2018
9-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
threshold is based on the range of possible combinations of TN and chlorophyll a and the
World Health Organization drinking water provisional guideline for microcystin of
1 (ig/L (WHO. 1998). The recommended threshold ranges to protect against frequency of
occurrence of high microcystin concentrations (0.1 MC < 1 (.ig/L) were 570 or
1,100 (ig/L TN paired with chlorophyll a concentrations of 37 and 3 (ig/L, respectively
(Yuan et al.. 2014). Decreasing the frequency of microcystin occurrence to 5% gave TN
concentrations of 250 or 400 (ig/L paired with chlorophyll a concentrations of 14 and
1 (.ig/L. In the most recent U.S. EPA National Aquatic Resource Surveys, microcystin
was detected in 12% of wetlands where water depth was sufficient to allow for
microcystin sampling (U.S. EPA. 20161) and in 39% of lakes (U.S. EPA. 2016h). The
U.S. EPA has recently provided human health advisories on allowable limits of 0.3 (ig/L
microcystins and 0.7 (ig/L cylindrospermopsin in drinking water over a 10-day exposure
period (U.S. EPA. 2015a. b).
Since the 2008 ISA, additional studies have indicated that the availability and form of N
influences algal bloom composition and toxicity, and inputs of inorganic N selectively
favor some HAB species. A recent bloom in Lake Okeechobee in Florida was dominated
by Microcystis, which depends on DIN for growth (Paerl and Scott. 2010). Multiyear
monitoring data from western Lake Erie showed that microcystin concentration peaks
coincided with inorganic N and that microcystin was significantly lower in years with
less inorganic N loading (Gobler et al.. 2016). In nutrient amendment experiments with
water collected from Lake Agawam in New York, abundances of toxic strains of
Microcystis were enhanced more than nontoxic strains by inorganic N, while nontoxic
strains responded to organic N more frequently than inorganic N (Davis et al.. 2010).
Microcystis populations were stimulated more frequently by N than by P in the assays.
Similarly, in a series of nutrient addition experiments using water from Sandusky Bay,
Lake Erie, bloom growth and microcystin concentrations responded more frequently to
the addition of inorganic and organic forms of N than to P addition, indicating that N
inputs may affect bloom size and toxicity (Davis et al.. 2015). Donald etal. (2011)
reported differential responses of phytoplankton to various forms of N in mesocosm
experiments in Wascana Lake, Saskatchewan. In this naturally P rich lake, addition of
NIL+ and urea promoted nonheterocystous cyanobacteria and algae, while increased
chlorophytes and some cyanobacteria were observed with NOs and urea. Microcystin
production increased with added N, although the response varied by the form of N and
the predominant algal taxon. In the same mesocosms, species-specific analyses indicated
144 individual taxathat exhibited distinct responses to N addition: 45 species showed
stimulated growth, 93 species had a limited response, and 6 species had suppressed
growth (Donald et al.. 2013).
9-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
9.2.6.2.
Enzymes
A recent study analyzed enzyme activity to characterize nutrient limitation in aquatic
systems. Microbial enzyme response to changes in N and P was found to vary in
terrestrial and aquatic compartments in Bear Brook watershed in Maine, the site of a
whole-watershed N enrichment experiment (Mincau et al.. 2014). Stronger effects were
typically observed in aquatic habitats. Although not explicitly the focus of the study, the
authors imply that P limitation is caused by N availability. Since 1989, ammonium
sulfate ([NH/J2SO4) has been applied bi-monthly to the watershed by helicopter,
(25.2 kg N/ha/yr). The activity of three enzymes (b-l,4-glucosidase [BG],
b-l,4-N-acetylglucosaminidase [NAG], acid phosphatase [AP]) in soil, leaf litter in both
terrestrial and stream habitats, and in stream biofilms were compared with the activities
of these enzymes in a reference watershed. In both terrestrial and aquatic habitats in the
Bear Brook watershed, BG and NAG activities were unaffected. Stream biofilms had
fivefold higher AP activity, while stream litter had eightfold higher AP activity. Increased
AP activity is indicative of enhanced P limitation in the treated watershed. Shorter
experimental P enrichments were used to characterize potential P limitation under
ambient and elevated N availability. In streams, the effects of acute P additions were
strongest with reduced AP activity and increased BG activity.
9.3. Community Composition, Species Richness, and Diversity
Increased N deposition to water bodies, either directly or via N leaching from soils,
increases available nutrients to aquatic biota and alters the balance of N and P. As
reported in the 2008 ISA, differences in resource requirements allow some species to gain
competitive advantage over others when nutrients are added, causing shifts in freshwater
community composition and structure as N concentration increases (Saros et al.. 2005;
Lafrancois et al.. 2004; Wolfe et al.. 2003). In eutrophic systems, water quality changes
associated with excess N nutrient inputs like lowered DO and algal blooms, which alter
habitat by covering up substrate, can also lead to declines in biodiversity (Hernandez et
al.. 2016). These pathways of N impacts have been identified as a threat for 50 aquatic
invertebrate species (mollusks) and 14 fish species that are listed, or candidates for
protection, under the U.S. Federal Endangered Species Act (Hernandez et al.. 2016). The
authors did not consider the sources of N in their analysis of biota affected by N
pollution.
Evidence for N effects on biodiversity in phytoplankton, zooplankton, and
macroinvertebrates in the 2008 ISA included observations from paleolimnological
studies, bioassays, and mesocosm and laboratory experiments. Community shifts and
June 2018
9-29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
decreased species richness of phytoplankton have been described in multiple studies
while fewer studies have considered zooplankton. The new literature described below
continues to report the effects of N enrichment on algal biodiversity and shows limited
evidence of effects at higher trophic levels (Table 9-3).
9.3.1. Archaea and Bacterial Diversity
The recent identification of ammonia-oxidizing archaea (AOA) and the classification of
these microbes in a distinct and novel phylum within Archaea (Thaumarchaeota) has led
to a large research effort to characterize their prevalence, community dynamics, and
ecological distribution (Schlcpcr and Nicol. 2010; Spang et al.. 2010). Spatio-temporal
dynamics and community structure of AOA appear to be affected by lake trophic status
and form of N (Mukhcrjcc et al.. 2016; Bollmann et al.. 2014; Berdieb et al.. 2013). In a
survey of nitrifying microbes in Lake Erie and Lake Superior, AOA outcompeted
ammonia-oxidizing bacteria (AOB) where ammonium concentrations are lower
(i.e., oligotrophic Lake Superior 0.287 (.iM NH4 versus. 2.45 (.iM in mesotrophic Lake
Erie) (Mukhcrjcc et al.. 2016).
9.3.2. Phytoplankton Diversity
Survey data, paleolimnological studies, and fertilization experiments in the 2008 ISA
reported species changes and reductions in plankton diversity in sensitive high-elevation
lakes in the western U.S. in response to increased availability of N (U.S. EPA. 2008a).
Available data reviewed in the 2008 ISA suggested that the increases in total N
deposition do not have to be large to elicit an ecological effect. Interlandi and Kilham
(2001) demonstrated that the highest species diversity was maintained at very low N
levels (<3 (.iM N) in lakes in the Yellowstone National Park region. A hindcasting
exercise determined that a change in Rocky Mountain National Park lake algae occurred
between 1850 and 1964 at only about 1.5 kg N/ha/yr wet N deposition (Baron. 2006).
Similar changes inferred from lake sediment cores of the Beartooth Mountains of
Wyoming also occurred at about 1.5 kg N/ha/yr deposition (Saros et al.. 2003).
Preindustrial inorganic N deposition is estimated to have been only 0.1 to 0.7 kg N/ha/yr
based on measurements from remote parts of the world (Holland et al.. 1999; Galloway et
al.. 1995). In the western U.S., preindustrial, or background, inorganic N deposition was
estimated by Holland et al. (1999) to range from 0.4 to 0.7 kg N/ha/yr.
June 2018
9-30
DRAFT: Do Not Cite or Quote

-------
Table 9-3 Summary of studies on nitrogen effects on species composition and biodiversity that have been
evaluated in the literature since the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria.
Region
Endpoint
Deposition
kg N/ha/yr
N Addition
Observation
Species
Ecosystem
Type
Reference
Isle Royale Phytoplankton 6-7 kg total 0, 5, 10, 20, or Chlorophytes had a reduced
National Park, community inorganic	40 [jg/L	response to DOM addition when N
Ml and Acadia structure	N/ha/yr wet	was added to the N limited lake
National Park.	deposition	(Sargent Lake in Isle Royale
ME	National Park). Increase in
abundance of diatoms (F.
crotonensis and Tabellaria
flocculosa) with DOM addition to
the lake.
Multiple
Boreal lakes
Daggett et
al. (2015)
Lake Tahoe,
CA and NV
Survey
Not specified None
Endemic benthic invertebrate taxa
have declined by 80 to 100%
since a survey of Lake Tahoe was
conducted in the 1960s. Changes
in the density and assemblage
structure of benthic invertebrates
mirrors increases in nutrient
enrichment and non-native
species in the lake. Macrophyte
occurrence has also declined.
Amphipods Stygobromus
lacicolus and S.
tahoensis declined
99.9%
No endemic turbellarians
but abundant in 1960s.
Endemic stonefly Capnia
lacustra decreased
93.5%
Endemic ostracods
Candona tahoensis
density decreased 83.4%
Subalpine,
oligotrophic
lake in
California
Caires et al.
(2013)
Ditch and Algal biofilm
Spread Creeks assemblage
in Grand Teton
National Park,
WY and Spring
Creek near
Wilson, WY
Not specified
0.5 M NaNOs;
0.5 M KH2PO4;
0.5 NaNOs, and
0.5 KH2PO4
Nutrient additions altered algal
biofilm assemblages in the
streams. NO3" addition inhibited
N2 fixer accumulation.
Phytoplankton
Streams
Kunza and
Hall (2013)
June 2018
9-31
DRAFT: Do Not Cite or Quote

-------
Table 9-3 (Continued): Summary of studies on nitrogen effects on species composition and biodiversity that have
been evaluated in the literature since the 2008 Integrated Science Assessment for Oxides
of Nitrogen and Sulfur-Ecological Criteria.
Region
Endpoint
Deposition
kg N/ha/yr
N Addition
Observation
Species
Ecosystem
Type
Reference
Front Range of
the Colorado
Rocky
Mountains
Phytoplankton
community
composition
Not specified
Added 930 pg/L
NO3" (with
background,
exposure was
1,240 pg NO3-;
added 93 pg/L
TDP)
Diatom abundance increased and
phytoplankton species
composition shifted in
nutrient-enriched mesocosms.
Principal component analysis
suggested that 21 % of the
variance in phytoplankton
community composition was
related to added nutrients, while
34% of the variance was due to
seasonal changes.
Phytoplankton
Alpine lake Gardner et
al. (2008).
12 alpine
ponds in Banff
National Park,
Canadian
Rockies
Phytoplankton
and periphyton
biomass/
zooplankton
biomass/
community
diversity
6.5-2.3 kg N/ha
/yr (2000-2010)
Highest levels
(30-90 kg N/ha/
yr) downwind of
urban and
agricultural
sources.
1 mg N/L
30 pg P/L
Simulating
deposition = 20 k
g/ha/yr
Periphyton outcompeted
phytoplankton for limiting
nutrients, indicating the
importance of considering both
benthic and pelagic primary
producers. High grazing pressure
by fairy shrimp affected results
predicted from chemical inference
and bioassay results regarding
effects of added nutrients on algal
communities.
Phytoplankton,
zooplankton, invertebrate
grazers, (fairy shrimp,
Anostraca: Branchinecta
paludosa)
Fishless,
nonglacial
ponds located
above tree
line
Vinebrooke
et al. (2014)
Banff	Phytoplankton
National Park,	abundance,
Canadian	effect on
Rockies	zooplankton
1,200 mg/L N
added to aquaria
with and without
sediment
A taxonomic shift from
omnivorous, raptorial-feeding
copepods to more effective
filter-feeding herbivorous daphnids
was observed with warming and N
addition. Warming and N
fertilization increased
phytoplankton abundance while
herbivory by Daphnia decreased
but only in the presence of
sediment. Effects of warming and
N differ both between trophic
levels and aquatic alpine habitats.
Zooplankton: herbivorous Alpine lake
Daphnia middendorffiana
and omnivorous
Hesperodiaptomus
arcticus
Thompson
et al. (2008)
June 2018
9-32
DRAFT: Do Not Cite or Quote

-------
Table 9-3 (Continued): Summary of studies on nitrogen effects on species composition and biodiversity that have
been evaluated in the literature since the 2008 Integrated Science Assessment for Oxides
of Nitrogen and Sulfur-Ecological Criteria.
Region
Endpoint
Deposition
kg N/ha/yr
N Addition
Observation
Species
Ecosystem
Type
Reference
Wapusk
National Park,
Canada
Phytoplankton
abundance,
effect on
zooplankton
Not specified
N (NH4NO3) and
P (KH2PO4) were
added to
increase the
nutrient
concentrations by
10x mean
ambient
concentrations
N limited lakes had statistically
significant different phytoplankton
community composition with more
chrysophytes and Anabaena spp.
compared to all other lakes.
41 phytoplankton taxa Subarctic
including	lakes and
Chlamydomonas spp., ponds
Sphaerocystis spp.,
Diatom a spp. and
Crugienella spp.
Svmons et
al. (2012)
Northern
Sweden
Phytoplankton
response to
whole lake
inorganic N
fertilization
<2	Dissolved KNO3
14 M N as KNOs
(in 2012) and 16
M HNOs (in
2013)
As DOC increased along a
gradient, community composition
shifted from nonflagellated toward
high DOC-adapted flagellated
autotrophs in the three fertilized
lakes. In the same set of lakes,
although phytoplankton biomass
increased, net zooplankton
responses were modest and
attributed by the authors to
incompatible stoichiometry of food
(phytoplankton) to consumers
(zooplankton)
Taxonomic groups:
chrysophytes,
crytophytes,
chlorophytes, diatoms,
dinoflagellates,
euglenophytes,
cyanobacteria,
picophytes.
Functional groups:
mixotrophic flagellates,
autotrophic flagellates,
nonflagellates,
heterotophic flagellates,
cyanobacteria,
picophytoplankton.
Boreal lakes
Deininqer
et al.
(2017a)
Northern
Sweden
Pelagic food
web response
to whole lake N
fertilization
<2	Dissolved
potassium nitrate
(14 M N as
KNOs) in 2012
and concentrated
nitric acid (14 M
N as HNOs) in
2013
Although phytoplankton biomass
increased, net zooplankton
responses were modest and
attributed to incompatible
stoichiometry of food
(phytoplankton) to consumers
(zooplankton). Therefore,
unconsumed phytoplankton could
accumulate in unproductive boreal
lakes with increased N deposition.
Crustacean zooplankton: Boreal lakes
calanoid copepods
(Eudiaptomus spp.),
cyclopoid copepods
(Cyclops spp.), and
cladocerans (mainly
Ceriodaphnia spp.,
Daphnia spp., Bosmina
spp., Diaphanosoma
brachyurum, Holopedium
gibberum, and Sida
spp.).
Deininqer
et al.
(2017b)
June 2018
9-33
DRAFT: Do Not Cite or Quote

-------
Table 9-3 (Continued): Summary of studies on nitrogen effects on species composition and biodiversity that have
been evaluated in the literature since the 2008 Integrated Science Assessment for Oxides
of Nitrogen and Sulfur-Ecological Criteria.
Region
Endpoint
Deposition
kg N/ha/yr
N Addition
Observation
Species
Ecosystem
Type
Reference
Denmark	Macroinverte- Not specified
brate
occurrence
None
Macroinvertebrate communities
did not change significantly with
TN based on an analysis of a
large number of concurrent
samples of macroinvertebrate
communities and chemical
indicators of eutrophication and
organic pollution (TP, TN,
NH4+-N, biological oxygen
demand [BOD5]) from 594 Danish
stream sites.
Plecopteran Leuctra spp. Streams
Isopod Asellus aquaticus
Dipteran Chironomus
spp.
Friberq et
al. (2010)
French Alps
Phytobenthos 7
species
richness,
grazing
pressure
Nutrient-diffusing A shift in taxonomic composition Diatoms
substrata
N treatment:
N = 113 [jg/N/hr
N + P:
N = 181 [jg/N/hr
P = 13 [jg/P/hr
Alpine lakes
of phytobenthos toward green
algae (less palatable to grazers)
from cyanobacteria and diatoms
was observed with N enrichment.
Benthic grazing by
macroinvertebrates was not
reduced.
Green algae
Cyanobacteria
Lepori and
Robin
(2014)
June 2018
9-34
DRAFT: Do Not Cite or Quote

-------
Table 9-3 (Continued): Summary of studies on nitrogen effects on species composition and biodiversity that have
been evaluated in the literature since the 2008 Integrated Science Assessment for Oxides
of Nitrogen and Sulfur-Ecological Criteria.
Region
Endpoint
Deposition
kg N/ha/yr
N Addition
Observation
Species
Ecosystem
Type
Reference
Bavaria,
Germany
Phytoplankton
community
composition
and
abundance
0, 1, 2, 8, 16, and Increasing N enrichment gradient Dominant phytoplankton Oligotrophic Poxleitner
32 mL of solution
of 30 g/L NOs"
and 10 mg/L
NH4+
affected phytoplankton
stoichiometry and community
composition and heterotrophic
nanoflagellate and ciliate
abundances, indicating N load
alters basal food web.
were Bacillariophyceae,
dinoflagellates,
Chrysophyceae, and
Cryptophyceae. Only a
few individuals of a few
Chlorophyceae and
Cyanophyceae (mainly
Anabaena spp.) were
observed.
Bacillariophyceae's most
abundant species were
Asterionella formosa,
Cyclotella spp., and
Fragilaria crotonensis.
Another abundant
Chrysophyceae species
Dinobryon divergens and
the dinoflagellate
Ceratium hirundinella.
lake
etal. (2016)
DOM = dissolved organic matter; KH2P04 = monopotassium phosphate; N = nitrogen; NaN03 = sodium nitrate; TN = total nitrogen; TDP = total dissolved phosphorus; TP = total
phosphorus.
June 2018
9-35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Some freshwater algae are particularly sensitive to the effects of added N and experience
shifts in species diversity and community composition with increased N deposition. A
general shift from chrysophytes that dominate lakes with low N to cyanophytes and
chlorophytes in lakes with higher N has been observed in studies reported in the
2008 ISA (Lafrancois et al.. 2004; Wolfe et al.. 2003; Jassbvetal.. 1994) and newer
studies (Svmons et al.. 2012; Pardo et al.. 2011a; Saros et al.. 2010). For example, in
phytoplankton sampling from 21 lakes and ponds in Wapusk National Park, Canada,
Svmons et al. (2012) observed that the N limited lakes had significantly different
phytoplankton community composition, with more chrysophytes and Anabaena spp.
compared to all other lakes.
Using data from the U.S. Geological Survey National Water Quality Assessment, Passv
(2008) assessed the responses of 2,426 benthic and 383 planktonic diatom communities
from 760 and 127 distinct localities, respectively, to nutrient limitation. As more
resources (basic cations, silica, iron, NH3, NO3 . and dissolved P) became limiting,
benthic diatom richness declined while phytoplankton richness increased.
9.3.2.1. Paleolimnological Studies
In the 2008 ISA, changes in diatom species assemblages, increases in cell numbers, and
pigment-inferred increases in whole-lake primary production were reported from
paleolimnological studies of mountain lakes that have only been disturbed by
atmospheric deposition and climate change. Sediment cores in oligotrophic lakes
receiving <5 kg N/ha/yr showed diatom community shifts that in part could be also
related to climate change. As reported in the 2008 ISA and Appendix 9.2.1 of this
appendix, two opportunistic species of diatom, A. formosa and F. crotonensis, now
dominate the flora of at least several lakes in the Rocky Mountains (Slcmmons etal..
2015; Arnett et al.. 2012; Saros et al.. 2010; Saros et al.. 2005; Saros et al.. 2003; Wolfe
et al.. 2003; Wolfe et al.. 2001; Interlandi and Kilham. 1998). In the southern Rocky
Mountains, this shift occurred in the 1950s in the south, with more recent shifts (1970s)
in the central region (Baron et al.. 201 lb). In most, but not all, of these studies, the
observed responses in phytoplankton were concordant with effects from increased
atmospheric N deposition. Increased abundance of A. formosa is also linked to changes in
lake temperature under conditions of declining N deposition and decreasing lake total N
concentration (Sivaraiah et al.. 2016).
Studies available since the 2008 ISA (summarized here and in Table 9-1) provide
additional evidence from historical records of lake sediments in assessing biological
changes associated with N enrichment over time. In a survey of 28 lakes in national parks
June 2018
9-36
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
across the western Great Lake region, 63% of study lakes experienced changes in the
diatom community correlated with lake sediment S15N (Hobbs et al.. 2016). About 36%
of sediment S15N was statistically correlated to some form of deposited N. Sediment
records from lakes in the Uinta Mountains, UT showed shifts in diatom community
composition and increasing abundance of the nitrophilous diatom A. formosa linked to
atmospheric deposition (Hundcv et al.. 2014). In sediment core sampling from
high-elevation lakes from national parks in Washington, Sheiblev et al. (2014) analyzed a
total of 56 sediment samples for diatom presence and abundance overtime. In the survey,
over 250 diatom species were identified; however, only Hoh Lake in Olympic National
Park showed clear evidence of impacts from N deposition based on changes in sediment
diatom assemblages and the presence of A. formosa, F. crotonensis, and Fragilaria
tenera. In high-elevation lakes in the Rocky Mountains with low atmospheric deposition
(<2 kg N/ha/yr) and low to moderate surface water NOs concentrations (<1 |_ig/L [below
detection] to 30 (ig/L). Arnett et al. (2012) observed that diatom assemblages were
already dominated by nitrophilous species like A. formosa or F. crotonensis. Because of
the abundance of species indicative of moderate N enrichment even at NO3
concentrations below the detection limit, it was not possible to identify a threshold of
reactive N response for diatom community change. In Baptiste Lake in Alberta, Canada,
a general shift in diatom assemblages away from N intolerant species and proliferation of
Stephanodiscus hantzschii, a nitrophilous diatom species, was observed (Adams et al..
2014). Paleolimnological analysis suggests that eutrophic conditions were present in the
lake for at least 150 years.
N deposition, along with climate change, was identified as a driver of diatom
compositional turnover (or beta diversity) in a synthesis of paleolimnological core
samples of 52 Arctic, alpine, and boreal montane lakes in North America and western
Greenland. Hobbs et al. (2010) stated that in all lakes, beta diversity was significantly
greater during the 20th century than the 19th century, with only a small and
nonsignificant difference in turnover between the 19th century and the 1550-1800
intervals (p = 0.86). Compared with forested montane boreal sites, both alpine and Arctic
lakes reveal greater diatom assemblage turnover in the 20th century. In Arctic lakes,
temperature appears to be the main factor affecting diatom composition turnover, while
in temperate lakes, N deposition appears to play a larger role. A meta-analysis of
200 lakes in Europe and North America showed diatom shifts in Arctic lakes around
1850 and temperate lakes around 1970 (Ruhland et al.. 2008). Observed changes were
primarily attributed to warming trends. The authors examined the diatom record from one
lake in detail (Whitefish Lake in Ontario) and reported that observations of diatom shifts
in this lake were not explained by N deposition but rather corresponded to temperature
changes.
9-37
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
New information on relative NOs inputs from glacial versus snowpack meltwaters
reported in Appendix 7 indicates water of glacial origin has higher levels of NO3, which
may influence the interpretation of biological data from high-elevation lakes and streams
in some regions of the U.S. (Slemmons et al.. 2015; Slemmons et al.. 2013; Saros et al..
2010; Baron et al.. 2009). In the central Rockies and Glacier National Park, fossil diatom
richness in snowpack-fed lakes was found to be higher (34 to 54 taxa) relative to lakes
fed by both glacial and snowpack meltwaters (12 to 26 taxa) (Saros et al.. 2010). In the
central Rockies, N deposition was 1.4 to 2.5 kg Nr/ha/yr, with 10 to 20% of the lakes
there receiving glacial meltwater. In Glacier National Park in the northern Rockies,
approximately 50% of lakes receive glacial meltwater and deposition was
2.0 to 3.4 kg Nr/ha/yr. In another study comparing sedimentary fossil diatom
assemblages in two adjacent lakes (one glacier-fed and the other snow-fed) in the central
Rocky Mountains, increased abundances of A. formosa, and F. crotonensis were
observed in the glacier-fed lake starting 1,000 years ago along with a decrease in diatom
species richness (Slemmons etal.. 2015). Shifts in the planktonic diatom communities
occurred after 1970 in snow-fed lakes (Slemmons et al.. 2017). These observations
suggest increased N inputs associated with glacial meltwater have altered the fossil algal
record and continue to affect algal communities in some glacier-fed lakes. No significant
differences in phytoplankton biomass and community structure were observed between
snowpack-fed and glacier and snowpack-fed lakes in the northern Cascade Mountains in
Washington in a comparison of data collected in 1989 and 2013 (Williams et al.. 2016b).
9.3.2.2. Bioassay, Mesocosm, and Laboratory Studies
Several experimental nutrient additions (mesocosm and bioassay studies) described in
Appendix 9.1.1.3 show that N limitation is common in freshwater systems (Elser et al..
2009b; Elser et al.. 2009a; Bergstrom et al.. 2008). Since the 2008 ISA, several N
addition studies have explored the effects of increased nutrients on phytoplankton
biomass and community structure (Daggett et al.. 2015; Gardner et al.. 2008). In the
Colorado Front Range, diatom abundance increased, and phytoplankton species
composition shifted in nutrient-enriched mesocosms (Gardner et al.. 2008). Principal
component analysis suggested that 21% of the variance in phytoplankton community
composition was related to added nutrients, while 34% of the variance was due to
seasonal changes. In a bioassay comparison of phytoplankton community structure
following N and DOM addition to a low DOC and N and P colimited water body (Jordan
Pond in Acadia National Park, ME) and an N limited lake (Sargent Lake in Isle Royale
National Park, MI), Daggett et al. (2015) observed increased chlorophytes following
DOM inputs in both lakes. In the N limited lake, an increase in the abundance of F.
June 2018
9-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
crotonensis and Tabellaria flocculosa was observed. There was a reduced response of
chlorophytes to DOM addition when N was added to the N limited lake.
Most N addition studies of phytoplankton communities have focused on N limited
systems. Poxleitner et al. (2016) conducted an N addition mesocosm experiment during
the spring season in a P limited oligotrophic lake in Upper Bavaria, Germany. An
increasing N enrichment gradient affected phytoplankton stoichiometry and community
composition. Only small effects of N enrichment were documented using the biovolume
of phytoplankton, the amount of particulate organic carbon, and the concentration of
chlorophyll a as indicators. There was an effect, however, on phytoplankton community
composition and heterotrophic nanoflagellate and ciliate abundances. Thus, changes in
food web dynamics were suggested for P limited lakes when N levels are increased.
9.3.2.3. Whole Lake Studies
Altered phytoplankton community composition and increased phytoplankton biomass
resulted frominorganic N fertilization of Reindeer Lake in northern Sweden (Jansson et
al.. 2001). In a whole-lake N fertilization study of three small boreal lakes in northern
Sweden, Deininger et al. (2017a) observed that changes in community composition of
phytoplankton were related to DOC rather than N addition. As DOC increased along a
gradient, community composition shifted from nonflagellated toward high DOC-adapted
flagellated autotrophs in the three fertilized lakes. In the same set of lakes, phytoplankton
biomass increased but net zooplankton responses were modest, attributed by the authors
to incompatible stoichiometry of food (phytoplankton) to consumers (zooplankton)
(Deininger et al.. 2017b).
9.3.3. Benthic Algal Diversity
Benthic algae are typically less sensitive than planktonic algae to nutrient inputs because
the former are associated with nutrient-rich substrates where sunlight reaches the
sediment surface (Spaulding et al.. 2015). Periphyton is typically more abundant than
phytoplankton in these shallow habitats (Vinebrooke et al.. 2014; Nvdick et al.. 2004).
Several recent studies have considered the effects of N on benthic algae characteristic of
shallow lakes and littoral zones where primary production is mainly controlled by the
availability of light (Vadeboncoeur et al.. 2008). In a stable isotope tracer study in Bull
Trout Lake, a subalpine lake in Idaho, the largest portion of added N was retained within
the periphyton (Epstein et al.. 2012). A directional change in benthic diatom species after
1960 that correlates with atmospheric deposition was observed in lake sediment cores
June 2018
9-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
from high-elevation shallow lakes in Grand Teton National Park (Spaulding et al.. 2015).
Liess et al. (2009) conducted a survey of periphyton along an N deposition gradient (2 to
12 kg N/ha/yr) in Sweden. Benthic algal composition in lakes in the northern part of the
country where N deposition was lower had a higher contribution from N fixing
cyanobacteria (5% of algal biovolume) than in the south. Overall, lakes were more N
limited in the North and P limited in the South. In another study using nutrient-diffusing
substrata in two small Alpine lakes in the Rhone Alps, France, N enriched substrata had a
greater phytobenthic biomass, and the taxonomic composition of the phytobenthos was
different between the treatment and control (Lcpori and Robin. 2014).
In pond fertilization experiments in 12 Ashless, nonglacial ponds in Snow Pass within the
Cascade Valley catchment of Banff National Park, Alberta, periphyton outcompeted
phytoplankton for limiting nutrients, indicating the importance of considering both
benthic and pelagic primary producers (Vincbrookc et al.. 2014). In these N limited
ponds, phytoplankton biomass increased significantly only when N was applied in the
absence of fairy shrimp. High grazing pressure by fairy shrimp appeared to suppress the
effects of added nutrients on algal communities, suggesting that trophic interactions are
important to consider to avoid missing the effects of N in alpine water bodies.
The algal assemblage response following 6 weeks of nutrient amendments in Wyoming
streams indicated that both N and P altered community structure of epilithic biofilm
(Kunza and Hall. 2013). Depending on which nutrient was added, N2 fixers reacted
differently than non-N2 fixers; an increase in non-N2 fixing diatoms in the N and N + P
additions was observed compared to N2 fixer dominance in control and P treatment.
Other studies consider how changes in benthic algal biodiversity affect ecosystem
processes such as chemical uptake. NO3 -N uptake differed among benthic algal
assemblages on transplanted rocks in a stream in Boise National Forest in central Idaho
(Baker et al.. 2009). Uptake of NO3 was highest in the green filamentous algae,
(dominated by the chlorophytes Spirogyra spp. and Rhizoclonium spp.), lowest in the
yellow patch type (dominated by the chlorophytes Spirogyra spp. and Bulbochaete spp.),
and intermediate in the brown patch type (dominated by diatoms, including Synedra spp.,
Cymbella spp., Fragilaria spp., and Epithemia spp.). NO3 —N uptake normalized to
chlorophyll a increased with algal composition and species richness in the three patch
types.
9.3.4. Zooplankton Diversity
Changes to aquatic food webs in response to N enrichment have not been as thoroughly
explored as changes to algal assemblages, but a few studies in the 2008 ISA showed
June 2018
9-40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
declines in zooplankton biomass (Lafrancois et al.. 2004; Paul et al.. 1995) in response to
N related shifts in phytoplankton biomass toward less palatable taxa with higher C:P
ratios (Elser et al.. 2001). New studies provide additional evidence for N deposition
effects on zooplankton through altered trophic interactions.
The nutritional status of zooplankton is influenced by the quantity and quality of food
items (Elser et al.. 2001). In lakes across an N deposition gradient in Norway,
zooplankton feeding on seston (organisms and nonliving matter in the water column)
were found to be affected by P limitation in phytoplankton (Elser etal.. 2010). The seston
in lakes with high N deposition had significantly higher C:P and N:P ratios due to
phytoplankton P limitation. Using an assay to measure alkaline phosphatase activity
(APA), which is overexpressed in animals with dietary P deficiency, Elseretal. (2010)
observed that the biomass-specific APA value differed considerably among taxa and that
N deposition was significantly associated with some species such as seston-feeding
zooplankton but not others, leading the authors to conclude that P limitation was
transferred up the food chain. In three P deficient lakes in Germany, N enrichment
experiments showed declines in mesozooplankton density and biomass, especially
cladocerans, with N additions in the range of projected increasing atmospheric deposition
(Trommcr et al.. 2017). An increase in seston C:P ratios was observed in one of the three
P deficient lakes in response to N enrichment.
A taxonomic shift from omnivorous, raptorial-feeding copepods to filter-feeding
herbivorous daphnids was observed with warming and N addition in a growth chamber
experiment (Thompson et al.. 2008). The study was conducted with water and sediment
collected from Pipit Lake in the Canadian Rockies in Banff National Park, Alberta. While
warming and N fertilization increased phytoplankton abundance, herbivory by Daphnia
middendorffiana decreased but only in the presence of sediment (pond conditions).
Findings demonstrated that the effects of warming and N differ both among trophic levels
and aquatic alpine habitats.
9.3.5. Macroinvertebrate Diversity
Macroinvertebrates such as aquatic insects, crustaceans, worms, and mollusks are
commonly sampled to assess water quality in federal and state monitoring programs (U.S.
EPA. 2016i). These organisms feed on algae and organic material and are an important
food source for fish and other wildlife. Few studies have considered the effects of
atmospheric N enrichment on biodiversity of freshwater macroinvertebrates. Due to the
existence of a survey of benthic invertebrates in Lake Tahoe, NV conducted in the 1960s,
it was possible to assess how the populations have changed with enrichment and the
June 2018
9-41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
introduction of non-native species. In this high alpine lake, atmospheric N contributions
are a significant portion (approximately 57%) of the total N loading to the lake (Sahoo et
al.. 2013). Caires et al. (2013) resurveyed the lake in 2008-2009 and observed declines of
80 to 100% in endemic benthic invertebrate taxa and changes to the community structure
of benthic invertebrate assemblages. Corresponding to these changes in lake biota, a
decrease in water clarity over the four decades between studies has been associated with a
shift in the bottom of the euphotic zone (1% light penetration) from 80 to 57 m (Chandra
et al.. 2005).
A recent study from Denmark suggests that macroinvertebrates may be poor indicators of
N enrichment. Friberg et al. (2010) measured stream macroinvertebrate occurrence along
gradients in organic pollution and eutrophication in 594 streams. They observed that the
occurrence of many taxa showed a stronger relationship to habitat condition than to
chemical variables. Overall, macroinvertebrate occurrence appeared to be related
primarily to biological oxygen demand, NH4+-N, and TP rather than TN. In a principal
component analysis using the United States Geological Survey (USGS) National
Ambient Water Quality Assessment (NAWQA) data set, Carlisle et al. (2007) identified
nutrients (NH/, NOs . TP) as a factor affecting macroinvertebrate occurrence, although
specific conductance, pH, and SO42 explained the greatest effects on macroinvertebrate
abundance. A 20% loss of macroinvertebrate taxa was indentified as a threshold for
degraded streams based on benthic macroinvertebrate sampling data from a subset of
basins in the eastern and midwestern U.S. (Carlisle and Meador. 2007).
9.3.6. Macrophyte Diversity
No U.S. studies of N effects on macrophyte (aquatic plant) community biodiversity in
atmospherically N enriched lakes and streams have been identified in the recent
literature, although declines in total macrophyte occurrence were noted in a resurvey of
Lake Tahoe that compared current samples with those from a lake survey conducted in
the 1960s (Caires et al.. 2013). Benthic invertebrate declines were most severe in
sampling sites where macrophytes were present in the 1960s but now absent, suggesting
that macroinvertebrate and macrophyte assemblages are closely associated in the lake.
Barker et al. (2008) conducted a mesocosm experiment using water and sediment from a
stream in Norfolk, U.K. Experimental tanks were planted with 11 macrophyte species
from the local environment. Constant P loadings were given to all tanks (50 |_ig P/L).
Nitrate loading varied from 1 to 10 mg NO3 -N/L. Macrophyte species richness
decreased with increasing N during the first year of treatment, and decreased in all
treatments above 1 mg NO3 -N/L during the second year. Barker et al. (2008) estimated
a threshold of 1.5 mg N/L for maintaining a stable species richness.
June 2018
9-42
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
9.3.7. Amphibian Diversity
No studies on N enrichment effects on amphibian diversity were reviewed in the
2008 ISA or identified in the current literature.
9.3.8. Fish Diversity
No studies of direct effects of N enrichment on freshwater fish diversity were reviewed in
the 2008 ISA. Post-2007 literature includes several behavioral endpoints in fish.
(Appendix 9.4.1).
9.4. Animal Behavior and Disease
In addition to changes in biological indicators (i.e., chlorophyll a, periphyton/microbial
biomass, diatoms, nutrient limitation shifts) and altered biodiversity, there is increasing
evidence for a role of N in behavior and disease in biota. These indirect responses may
impact the fitness of organisms inhabiting nutrient-enriched waters.
9.4.1. Behavior
Nutrient enrichment of freshwater systems has recently been shown to alter behavioral
endpoints in an invertebrate, an amphibian, and a fish species in laboratory exposures.
N03 exposure (21.4, 44.9, 81.8, 156.1 mg NO, /L) was shown to decrease the velocity
of movement in the aquatic snail Potamopyrgus antipodarum (Alonso and Camargo.
2013). Reproductive impairments (decreased number of newborns) were observed at all
tested concentrations. The NOs concentrations used in this study are much higher than
typically measured in remote freshwater catchments affected by atmospheric deposition.
In the presence of chemical cues of the predator nymphs of the dragonfly (Anax
imperator), western spadefoot toad (Pelobates cultripes) tadpoles typically decrease
swimming activity by 44% (Polo-Cavia et al.. 2016). When predator chemical cues were
added to water containing NH4NO3 tadpoles did not alter swimming activity. The
concentrations of NH4NO3 used in the study (20 mg/L and 80 mg/L) were not lethal to
the tadpoles, and no altered swimming activity was observed with the added N except in
the presence of predator chemical cues.
In the three-spined stickleback Gasterosteus aculeatus, a fish that inhabits both
freshwater and brackish habitats, changes to water quality associated with eutrophication
June 2018
9-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
(i.e., turbidity associated with algal blooms) have impacted social and reproductive
behaviors in laboratory studies. These studies are reviewed in Appendix 10.
9.4.2. Disease
The interactions of N enrichment and disease in biota is a relatively new area of research
in N impacted freshwater systems. The interactions characterized to date are complex and
involve indirect effects on host-parasite interactions (Johnson et al.. 2010; Johnson ct al..
2007). A host-parasite relationship potentially impacted by increased N to aquatic
systems is that of the fungal pathogen Metschnikowia bicuspidata, which is parasitic to
crustacean zooplankton Daphnia dentifera (Dallas and Drake. 2014). In a series of
bioassays designed to assess the effects of N on host and pathogen, D. dentifera were
exposed to six NOs concentrations (0.4, 2, 4, 8, 16, and 32 mg NOs /L) and then
inoculated withM bicuspidata. NOs, decreased D. dentifera population size and
increased infection prevalence. Next, ambient levels of N (0.4 mg NO3 -N/L) and a
concentration representative of a moderately polluted system (12 mg NO, -N/L) were
used to test the influence of NO3 on pathogen dose, infection prevalence, and host
fitness (growth and fecundity). No effects were observed on the growth rate of D.
dentifera; however, greater infection prevalence was associated with increased NO, . and
in general, both host fecundity and infection intensity decreased with increasing pathogen
dose.
9.5. Summary of Thresholds, Levels of Deposition at Which
Effects Are Manifested, and Critical Loads
Since the 2008 ISA, additional thresholds of response to N have been identified that are
useful for critical load development in water bodies impacted by N nutrient effects.
Critical loads for N nutrient enrichment in U.S. freshwater ecosystems are summarized in
Table 9-4. Factors contributing to uncertainty in N deposition estimates for assessment of
critical loads according to Pardo etal. (2011a) include "(1) the difficulty of quantifying
dry deposition of nitrogenous gases and particles to complex surfaces; (2) sparse data,
particularly for arid, highly heterogeneous terrain (e.g., mountains); and (3) sites with
high snowfall or high cloud water/fog deposition, where N deposition tends to be
underestimated." N deposition estimates at high-elevation sites such as those in the
Rocky and Sierra Nevada mountains are associated with considerable uncertainty,
especially uncertainty for estimates of dry deposition (Appendix 2). For sensitive
receptors such as phytoplankton shifts in high-elevation lakes, N deposition model bias
may be close to, or exceed, predicted critical load values (Williams et al.. 2017a).
June 2018
9-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
Physical, chemical, and ecological variability across lakes affect their response to N
deposition and contribute to uncertainty of critical load estimates (Appendix 9.1.1.2V A
review by Bowman et al. (2014) noted that current N critical loads for sensitive alpine
systems may not be protective under future climate scenarios of warmer summer
temperatures and a shorter duration of snow cover.
Available data from the 2008 ISA suggest that the increases in total N deposition do not
have to be large to elicit an ecological effect in remote alpine lakes. A hindcasting
exercise determined that the shift in Rocky Mountain National Park lake algae
composition that occurred between 1850 and 1964 was associated with wet N deposition
that was only about 1.5 kg N/ha/yr (Baron. 2006). Similar changes inferred from lake
sediment cores of the Beartooth Mountains of Wyoming occurred at about
1.5 kg TN/ha/yr deposition (Saros et al.. 2003).
9-45
DRAFT: Do Not Cite or Quote

-------
Table 9-4 Summary of critical loads for nitrogen eutrophication for surface
water in the U.S. [adapted from Pardo et al. (2011c) with newer
studies added].
Ecosystem
Site
Critical Load for
N Deposition
kg N/ha/yr
Comments
Reference/
HERO ID
Alpine stream
Southern Rockies/Loch
Vale Rocky Mtn. Nat.
Park
2
Modeled
Baron et al.
(1994)
Alpine lake
Northern
Rockies/Beartooth Mtns.,
WY
1.5
Paleolimnological, shifts in
diatom assemblages with
Fragilaria crotonensis and
Cyclotella bodanica
increasing to comprise
approximately 30% each
of total assemblage
Saros et al.
(2003)
Alpine lakes
Rocky Mountains
2.5
Surveys and references
therein
Berqstrom and
Jansson(2006)
Alpine lakes
Eastern Sierra Nevada
and Greater Yellowstone
Ecosystem
1.4
Paleolimnological,
increase in the relative
abundances of
Asterionella formosa and
Fragilaria crotonensis
Saros et al.
(2011)
High-elevation
lakes of western
and eastern
U.S.
Rocky Mountains, Sierra
Nevada, Northeast
1.0 to 3.0 (western
lakes)
3.5 to 6.0
(northeastern lakes)
Based on N deposition
calculated in three
different ways and on lake
NO3" concentrations
Baron et al.
(2011b)
Remote lakes

2.0 (western lakes)
8.0 (eastern lakes)
Based on value of N
deposition at which
significant NO3" leaching
begins to occur
Pardo et al.
(2011c)
Alpine lakes
Rocky Mountain National
Park
>1.5
Based on NO3" threshold
of 0.5 |jM (concentration
which elicits a growth
effect in the diatom
A. formosa)
Nanus et al.
(2012)
Alpine lake
Hoh Lake, Olympic
National Park, WA
1.0-1.2
Increased relative
abundances of
Asterionella formosa and
Fragilaria tenera
Diatom assemblage shift
Sheiblev et al.
(2014)
Alpine lakes
Greater Yellowstone area
<1.5 to >4.0a
Modeled; based on
threshold value of NO3"
(0.5 to 2.0 peq/L)
Nanus et al.
(2017)
June 2018
9-46
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Table 9-4 (Continued): Summary of critical loads for nitrogen eutrophication for
surface water in the U.S. [adapted from Pardo et al. (2011c)
with newer studies added].
Ecosystem
Site
Critical Load for
N Deposition
kg N/ha/yr
Comments
Reference/
HERO ID
Western U.S.
lakes
Remote mountain lakes
across the western U.S.
4.1a
Based on phytoplankton
biomass nutrient limitation
shifts
Williams et al.
(2017b)
Western U.S.
lakes
Remote mountain lakes
across the western U.S.
2.0a
Modeled to reduce
occurrence of false
negatives to near zero
Williams et al.
(2017b)
N = nitrogen; N03 = nitrate.
atotal N
Since the release of the 2008 ISA, work has continued on identifying thresholds of
response to N deposition that can be used to calculate critical loads in sensitive
freshwater systems. Threshold values for phytoplankton biomass growth have been
identified for nine western mountain lakes in North Cascades, Mount Ranier, and
Olympic national parks (13 to 25 (.ig DIN/L) (Williams et al.. 2016a) and in lakes in the
Sierra Nevada [0.3 to 4 (.iniol N/L (5 to 56 (.ig N/L); (Heard and Sickman. 2016)1.
Williams et al. (2017b) identified nutrient limitation shift thresholds (21 to 53 |_ig N/L)
for remote western U.S. mountain lakes. A nutrient threshold for surface water NOs for
growth of the diatom A. formosa was developed for high-elevation lakes in the Rocky
Mountains where N deposition is prevalent (Nanus et al.. 2012). The NO3 threshold of
0.5 (imol/L (31 |ig/L) was then used to estimate areas in Rocky Mountain National Park
that exceed critical load values.
Additional critical loads for nutrient enrichment of freshwaters developed since the
2008 ISA include those in the western and northeastern U.S. Pardo etal. (2011c)
estimated a critical load of 2.0 kg N/ha/yr for western lakes and 8.0 kg N/ha/yr for eastern
lakes based on the value of N deposition at which significant NO3 leaching begins to
occur. Using data from 125 lakes in the Greater Yellowstone area, including sensitive
lakes in both Grand Teton and Yellowstone national parks, Nanus et al. (2017) estimated
aquatic critical load values and identified locations with predicted critical load
exceedances based on a threshold value of NO3 concentration in lake water selected as
indicative of biological harm (0.5 to 2.0 j^ieq/L in this study). Critical loads of TN
deposition ranged from <1.5 ± 1.0 kg N/ha/yr to >4.0 ± 1.0 kg N/ha/yr. Exceedance
estimates were as high as 48% of the Greater Yellowstone area study region.
Baron etal. (2011b) found that in the western high-elevation lakes, increased primary
productivity and changes to algal diversity can occur with only minimal inputs of N
June 2018
9-47
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
deposition. They estimated that the thresholds, or critical loads, for nutrient enrichment
are 1.0 to 3.0 kg N/ha/yr for the western mountains (Sierra Nevada and Rocky
mountains) (Table 9-5). For minimally disturbed lakes in the Northeast, a critical load of
3.5 to 6.0 kg N/ha/yr was estimated, but independent biological measures for nutrient
enrichment are lacking in this region. In another study from the eastern Sierra Nevada
and Greater Yellowstone Ecosystem, Saros etal. (2011) determined a critical load of
1.4 kg N/ha/yr wet N deposition by modeling wet deposition rates from the period in
which diatom shifts first occurred. The shifts were identified from sediment cores
between 1960 and 1965 in the eastern Sierra Nevada, and starting in 1980 in the Greater
Yellowstone Ecosystem. Using core samples, Sheiblev et al. (2014) identified diatom
changes corresponding to N inputs from 1969-1975 in Hoh Lake in Washington and
established a critical load of 1.0-1.2 ± 0.01 kg N/ha/yr for the lake.
Using phytoplankton biomass N to P limitation shifts as the basis for critical load
calculations, Williams et al. (2017b) determined an empirical critical load of
4.1 kg/TN/ha/yr for remote high-elevation lakes across the western U.S. The critical
loads were calculated as the total (wet + dry) N deposition rate at which point below the
critical load, N limitation is more likely than P limitation and above the critical load, P
limitation is more likely than N limitation. Modeled critical loads using DIN:TP and DIN
response categories yielded an average critical load of 3.8 kg/TN/ha/yr for the lakes.
Modeled critical loads ranged from 2.8 to 5.2 kg/TN/ha/yr and correctly predicted
exceedances in 69% of lakes using NOs -N. The authors conducted a performance
evaluation using the NO3 -N univariate model and identified a critical load of
2.0 kg/TN/ha/yr to reduce the likelihood of false negatives to near zero.
9-48
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Table 9-5 Summary of mean lake nitrate (NO3-) concentrations, inorganic
nitrogen deposition amounts, and nutrient enrichment inflection
points where lake NO3- concentrations reflect increased nitrogen
deposition, [from Baron et al. (2011b)1.
Region
Mean
Lake
NOs"
Mean
1997-2006
NADP N
Deposition
Nutrient
Enrichment
Inflection
Point
(NADP)
Mean 1997-2006
(PRISM + NADP)
Deposition
Nutrient
Enrichment
Inflection Point
(PRISM + NADP)
2002 Total
N
Deposition
Nutrient
Enrichment
Inflection
Point
(Total N)
Sierra
Nevada
(n = 30)
2.7
(1.10)
1.5 (0.22)
1.5
2.5 (0.34)
1.5
3.4 (0.47)
2.0
Rocky
Mountains
(n = 285)
3.7
(1.90)
1.2 (0.27)
1.0
3.0 (0.28)
2.0
2.0 (0.24)
3.0
Northeast
(n = 216)
14.4
(0.76)
4.7 (0.18)
3.5
5.2 (0.18)
3.5
3.5 (0.20)
6.0
N = nitrogen; NADP = National Atmospheric Deposition Program; N03 = nitrate; PRISM = Parameter-Elevation Regressions on
Independent Slopes Model; PRISM + NADP = concentrations from NADP with PRISM precipitation values.
Note: Inorganic N deposition was calculated three ways, as was described in Baron et al. (2011b). Coefficients of variation are in
parentheses.
Mean lake N03" (in micromoles per liter), inorganic nitrogen (N) deposition amounts (in kg N/ha/yr), and nutrient enrichment
inflection points where lake N03" concentrations increase in response to increasing N deposition.
Source: Aber et al. (2003). Sickman et al. (2002). U.S. Forest Service's Air Resource Management online database
(www.fs.fed.us/ARMdata).
9.6. Summary and Causal Determination
In the 2008 ISA, the body of evidence was sufficient to infer a causal relationship
between N deposition and the alteration of species richness, community composition, and
biodiversity in freshwater ecosystems. New evidence from 2008 to the present, from
paleolimnological surveys, fertilization experiments, gradient studies, phytoplankton
community responses, and indices of biodiversity continue to show effects of N loading
to sensitive freshwater systems. As reported in the 2008 ISA, freshwater systems in
which N has been observed to influence ecological processes either received extremely
high inputs [e.g., (Dumont et al.. 2005)1. or had very low initial N concentrations and
responded rapidly to the additional inputs (Bcrgstrom and Jansson. 2006; Baron et al..
2000). As summarized in the 2008 ISA and supported by data in newer studies, nutrient
enrichment effects on freshwater ecosystems from atmospheric deposition of N are most
likely to occur in lakes and streams that have low primary productivity and low nutrient
levels and that are located in the most undisturbed areas with no local pollution sources.
June 2018
9-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Even small inputs of N in these water bodies can increase nutrient availability or alter the
balance of N and P, which can stimulate growth of phytoplankton. As reported in the
2008 ISA and further strengthened with new evidence in this review, there is consistent
and coherent evidence for species composition changes and reductions in diversity,
especially for primary producers from high-elevation lakes in the western U.S., in
response to increased availability of N. New information is consistent with the
conclusions of the 2008 ISA that the body of evidence is sufficient to infer a causal
relationship between N deposition and changes in biota, including altered growth
and productivity, species richness, community composition, and biodiversity due to
N enrichment in freshwater ecosystems.
New data have not appreciably changed the consistent and coherent evidence at the time
of the 2008 ISA that freshwater systems that are likely to be most impacted by nutrient
enrichment due to atmospheric deposition of N are remote, oligotrophic, high-elevation
water bodies with limited local nutrient sources and with low N retention capacity. N
inputs to these ecosystems are linked to changes in biota, especially increased algal
growth and shifts in algal communities. Freshwater systems sensitive to N nutrient
enrichment include those in the Snowy Range in Wyoming, the Sierra Nevada
Mountains, and the Colorado Front Range. A portion of these lakes and streams where
effects are observed are in Class I wilderness areas (Williams et al.. 2017b; Clow et al..
2015; Nanus et al.. 2012).
Recent research further supports the 2008 ISA findings that N limitation is common in
oligotrophic waters in the western U.S. (Elser et al.. 2009b; Elser et al.. 2009a). Shifts in
nutrient limitation, from N limitation, to between N and P limitation, or to P limitation,
were reported in some alpine lake studies reviewed in the 2008 ISA and in this review.
Since the 2008 ISA, several meta-analyses have reported an increase in P deposition to
water bodies (Stoddard et al.. 2016; Brahnev et al.. 2015; Tipping et al.. 2014) and
highlight the need to account for how sustained P deposition can modify the effects of
anthropogenically emitted N deposition on productivity. P addition delayed the shift to P
limitation (prolonged N limitation) for phytoplankton.
The body of evidence for biological effects of N enrichment and altered N:P ratios in
remote freshwater systems (where atmospheric deposition is the predominant source of
N) is greatest for phytoplankton, the base of the freshwater food web. Lake surveys,
fertilization experiments, and nutrient bioassays reported in the 2008 ISA and in this
review show increased pelagic and benthic algal productivity (as indicated by
chlorophyll a concentration) to be strongly related to increased N concentration. An
increase in lake phytoplankton biomass with increasing N deposition was reported in
several regions, including the Snowy Range in Wyoming and across Europe. New studies
9-50
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
in the Rocky Mountains of Colorado where atmospheric deposition ranged from 2 to
7 kg N/ha/yr support observations from the 2008 ISA which showed correlations between
greater chlorophyll a response and higher rates of deposition (Elser et al.. 2009a'). Several
recent studies have considered the effects of N on benthic algae characteristic of shallow
lakes and littoral zones where availability of light is the main factor controlling primary
production. A few studies have shown that periphyton outcompeted phytoplankton for
limiting nutrients, indicating the importance of considering both benthic and suspended
primary producers.
Changes in phytoplankton species diversity and community structure reported in the
2008 ISA (Lafrancois et al.. 2004; Wolfe et al.. 2003; Jassbvetal.. 1994) and newer
studies (Svmons et al.. 2012; Saros et al.. 2010) show a general shift from chrysophytes
that dominate lakes with low N to cyanophytes and chlorophytes in lakes with higher N.
In the 2008 ISA, diatom assemblage shifts were reported from the literature at total N
deposition as low as 1.5 kg N/ha/yr (Saros et al.. 2003). This evidence was further
supported by a hindcasting exercise that determined the change in Rocky Mountain
National Park lake algae that occurred between 1850 and 1964 was associated with an
increase in wet N deposition that was only about 1.5 kg N/ha/yr (Baron. 2006). Two
nitrophilous species of diatoms, A. formosa and F. crotonensis, are dominant in lakes
with higher N and serve as biological indicators ofN enrichment; however, increased
relative abundance of A. formosa has also been attributed to lake warming in some
regions where N deposition is decreasing (Sivaraiah et al.. 2016). Some shifts in algal
biodiversity observed in high-elevation waters are attributed to climate change or nutrient
effects and climate as costressors (Appendix 13).
The role of N in freshwater HAB formation has been further researched since the
2008 ISA. Additional evidence continues to show that the availability and form ofN
influences algal bloom composition and toxicity and that inputs of inorganic N
selectively favor some HAB species, including those that produce microcystin.
Microcystin is prevalent in U.S. waters as reported in recent regional and national
surveys. Although the risk of HAB formation is low in high-elevation oligotrophic water
bodies where N deposition is the dominant source ofN, transport of atmospheric inputs
can exacerbate eutrophic conditions in downstream water bodies. Increased
understanding of the role of N as a limiting nutrient in many freshwater systems has led
to recommendations to consider both N and P in nutrient reduction strategies (Dodds and
Smith. 2016; Gobler et al.. 2016; Paerl et al.. 2016b; Lewis etal.. 2011; Scott and
McCarthy. 2010; Conlev et al.. 2009; Paerl. 2009; Lewis and Wurtsbaugh. 2008).
Since the 2008 ISA, empirical and modeled critical loads for the U.S. have been
estimated based on surface water NOs concentration, diatom community shifts, and
9-51
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
phytoplankton biomass growth nutrient limitation shifts. A critical load ranging from 3.5
to 6.0 kg N/ha/yr was identified for high-elevation lakes in the eastern U.S. based on the
nutrient enrichment inflection point [where NOs concentrations increase in response to
increasing N deposition; (Baron et al.. 201 lb)I. Another critical load of 8.0 kg N/ha/yr for
eastern lakes based on the value of N deposition at which significant increases in surface
water NO3 concentrations occur was estimated by Pardo etal. (2011c). In both Grand
Teton and Yellowstone national parks, critical loads for total N deposition ranged from
<1.5 ± 1.0 kg N/ha/yr to >4.0 ± 1.0 kg N/ha/yr (Nanus et al.. 2017). Exceedance
estimates were as high as 48% of the Greater Yellowstone area study region, depending
on the threshold value of NO;, concentration in lake water selected as indicative of
biological harm.
Additional critical loads have been identified since the 2008 ISA for eastern Sierra
Nevada lakes, Rocky Mountain lakes, the Greater Yellowstone Ecosystem, and Hoh
Lake, Olympic National Park ITable 9-4; (Nanus et al.. 2017; Sheiblev et al.. 2014; Pardo
et al.. 2011c; Saros et al.. 2011)1. The identified values fall near or within the range of 1.0
to 3.0 kg N/ha/yr for western lakes (Baron et al.. 201 lb). An empirical critical load of
4.1 kg/TN/ha/yr above which phytoplankton biomass P limitation is more likely than N
limitation was identified by Williams et al. (2017b) for the western U.S. Modeled critical
loads ranged from 2.8 to 5.2 kg/TN/ha/yr, and a performance analysis indicated that a
critical load of 2.0 kg/TN/ha/yr would likely reduce the occurrence of false negatives to
near zero.
The evidence for N effects on other freshwater biota is not as extensive as for
phytoplankton. Since the 2008 ISA, new research on archaea and bacterial diversity in
freshwater systems suggests that these organisms respond to lake trophic state and the
form ofN present. At higher trophic levels, zooplankton responses to N inputs are
attributed to changes in food quality which can potentially alter food web interactions
(Deininger et al.. 2017b; Meunier et al.. 2016; Elser et al.. 2009a). A few studies in the
2008 ISA showed declines in zooplankton biomass (Lafrancois et al.. 2004; Paul et al..
1995) in response to N related shifts in phytoplankton biomass toward less palatable taxa
with higher C:P ratios (Elser et al.. 2001). Limited studies since the 2008 ISA suggest
that atmospheric N inputs are linked to taxonomic shifts and declines in some
invertebrates, although the effects attributed to N are difficult to separate from other
stressors such as climate change and invasive species. Water quality changes associated
with excess N nutrient inputs, such as lowered DO and algal blooms that alter habitat by
covering up substrate, can also lead to declines in biodiversity in macroinvertebrates and
fish including species listed under the federal Endangered Species Act (Hernandez et al..
2016). Although little research has been conducted in freshwater systems, some evidence
suggests that increased turbidity associated with algal blooms may affect reproductive
9-52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
and social behaviors in fish (Appendix 10). Emerging research on disease in biota suggest
that N enrichment may modify relationships such as host susceptibility to parasites and
pathogens.
Since the 2008 ISA, further studies have shown that both trophic interactions and DOC
modify ecosystem response to N loading. In a whole-lake N fertilization study, Deininger
et al. (2017a) observed that changes in community composition of phytoplankton were
related to DOC rather than N addition to small N limited boreal lakes. As DOC increased
along a gradient, community composition shifted from nonflagellated toward high
DOC-adapted flagellated autotrophs in the three fertilized lakes. In the same set of lakes,
although phytoplankton biomass increased, net zooplankton responses were modest and
attributed by the authors to incompatible stoichiometry of food (phytoplankton) to
consumers (zooplankton) (Deininger et al.. 2017b). A study in Banff National Park,
Canada indicated that grazing pressure on algae may have negated the effects of nutrient
inputs on primary producers (Vinebrooke et al.. 2014). Limited evidence suggests a
decreased rate of leaf litter decomposition by microbial communities with N loading to
streams.
New studies show that the N contribution from glacial meltwater (which has higher NOs
relative to water from melting snow) affects diatom community composition in
high-elevation lakes and streams in some regions of the U.S. (Slemmons et al.. 2015;
Slemmons et al.. 2013; Saros et al.. 2010). N deposition to snow and glaciers are
important sources of N to alpine lakes and streams that are fed by meltwaters. In a
comparison of biological response in snowpack-fed lakes versus lakes with both glacial
and snowpack meltwater in the Rockies, fossil diatom richness was higher in the
snowpack-fed lakes (34 to 54 taxa) compared to lakes with both glacial and snow
meltwater inputs [12 to 26 taxa; (Saros et al.. 2010)1. Characterizing diatom community
responses to meltwater sources is important for interpreting the biological effects of N
deposition in high-elevation systems with both glacial and snow inputs.
9-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
APPENDIX 10. BIOLOGICAL EFFECTS OF
NITROGEN ENRICHMENT IN
ESTUARIES AND NEAR-COASTAL
SYSTEMS
This appendix characterizes the biological effects of nitrogen (N) enrichment in estuaries
(areas where freshwater from rivers meets the salt water of oceans), coastal lagoons, coral
reef ecosystems, and open ocean areas near coastlines. An overview of N inputs to
coastal systems, including characteristics and identification of areas of the U.S. sensitive
to nutrient over-enrichment (Appendix 10.1) is followed by discussions of the indicators
of nutrient enrichment (Appendix 10.2). its effects on biodiversity and ecosystem
structure and function (Appendix 10.3). animal behavior, and disease (Appendix 10.4).
and the role of N in nutrient enhanced coastal acidification (Appendix 10.5).
Appendix 10.6 summarizes the thresholds of biological effects ofN in coastal regions,
and Appendix 10.7 reviews the causal determinations based on a synthesis of new
information and previous evidence from prior N assessments.
10.1. Introduction
In the 2008 Integrated Science Assessment for Oxides of Nitrogen and Sulfur-Ecological
Criteria (2008 ISA), the body of evidence was sufficient to infer a causal relationship
between N deposition and the alteration of species richness, species composition, and
biodiversity in estuarine systems (U.S. EPA. 2008a). Nitrogen pollution is the major
cause of harm to the majority of estuaries in the U.S. (Bricker et al.. 2008; NRC. 2000).
In estuaries, N from atmospheric and nonatmospheric sources contributes to increased
primary productivity, leading to eutrophication (Figure 10-1). Estuary eutrophication is a
process of increasing nutrient over-enrichment indicated by water quality deterioration,
resulting in numerous adverse effects including areas of low dissolved oxygen (DO)
concentration (hypoxic zones), species mortality, and harmful algal blooms (HABs).
Eutrophic systems are characterized by an increase in the rate of supply of organic matter
(primary production and organic carbon accumulation) in excess of what an ecosystem is
normally adapted to processing (Diaz et al.. 2013; Nixon. 1995). In the 2008 ISA the
strongest evidence for a causal relationship was from changes in biological indicators of
nutrient enrichment including chlorophyll a, macroalgal "seaweed" abundance, HABs,
DO, and submerged aquatic vegetation (SAV). Biological effects of increasing nutrient
enrichment also include changes in biodiversity in these systems (Chapter 1.2.2.4). For
this ISA, new information is consistent with the 2008 ISA and the causal determination
June 2018
10-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
has been updated to reflect more specific categories of effects. The body of evidence is
sufficient to infer a causal relationship between N deposition and changes in biota,
including altered growth, total primary production, total algal community biomass,
species richness, community composition, and biodiversity due to N enrichment in
estuarine environments.
Since the 2008 ISA, N loading has been recognized as a possible contributing factor to
acidification of coastal waters (Appendix 10.5). Dissolution of atmospheric
anthropogenic CO2 into the ocean has led to long-term decreases in pH. Increased
production of CO2 from microbial degradation of organic matter associated with
eutrophication and respiration of algae and seagrasses during the night along with
atmospheric anthropogenic CO2 inputs can make the ocean water more acidic (Wilson et
al.. 2016; Wallace et al.. 2014; Wetz and Yoskowitz. 2013; Sunda and Cai. 2012; Cai et
al.. 2011c; Howarth et al.. 201IV Organisms that produce calcium carbonate shells such
as calcareous plankton, oysters, clams, sea urchins, and coral may be affected by
long-term decreases in pH (Mostofa et al.. 2016; Pfister et al.. 2014; Kroeker et al.. 2013;
Sunda and Cai. 2012). The body of evidence is suggestive, but not sufficient to infer, a
causal relationship between N deposition and changes in biota, including altered
physiology, species richness, community composition, and biodiversity due to
nutrient-enhanced coastal acidification.
In the previous ISA, there was a strong scientific consensus that N is the principal cause
of coastal eutrophication in the U.S. (U.S. EPA. 2008a; NRC. 2000). On average, human
activity has likely contributed to a sixfold increase in the N flux to the coastal waters of
the U.S. over the past several decades, and excessive N represents the most significant
coastal pollution problem (Bricker et al.. 2008; Bricker et al.. 2007; Howarth and Marino.
2006; Bricker etal.. 2003; Howarth et al.. 2002; NRC. 2000). Many coastal areas receive
high enough levels of N input from human activities to cause eutrophication (Bricker et
al.. 2007; Howarth et al.. 1996a; Vitousek and Howarth. 1991). As reported in the 2008
ISA, N driven eutrophication of shallow estuaries has increased over the past several
decades, and environmental degradation of coastal ecosystems is now a widespread
occurrence (Bricker et al.. 2007; Paerl et al.. 2000).
10-2
DRAFT: Do Not Cite or Quote

-------
Atmospheric Deposition
iJvAV
~ - - •-«
% ** j Phytoplankton Bloom
• thrives on nutrients
-o
Dissolved Oxygen
from wave action
and photosynthesis
1 1"
I
4
~
~ 4 "
Dead ~
materials
#
settles m
Dissolved Oxygen
trapped in the upper,
lower-salinity layer
Decomposition
*
. Dissolved Oxygen used up
1 by microorganism respiration
1 i»				.
* * i	Nutrients
'» released by bottom sediments
Lower-density
surface water
Higher-density
bottom water
Dissolved Oxygen consumed
Ir~
Fish will avoid
hypoxia if possible
Shellfish
and other
benthic
organisms
unable
to escape
hypoxia
Decomposition of organic
matter in sediments
N = nitrogen.
Notes: atmospheric N deposition is one of the sources of nutrient enrichment to coastal areas.
Source: modified from U.S. EPA (2012b).
Figure 10-1 Eutrophication can occur when the availability of nutrients
increases above normal levels.
June 2018
10-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Coastal systems are linked to terrestrial N processes along the freshwater to ocean
continuum as nutrients deposited to the watershed move downstream. As described in the
2008 ISA, in coastal ecosystems, the nutrients most commonly associated with
phytoplankton growth are N, phosphorus (P), and silicon (Si). The relative proportions of
these nutrients are important determinants of primary production, food web structure, and
energy flow through the ecosystem (Turner et al.. 1998; Justic et al.. 1995b; Justic et al..
1995a; Dortch and Whitledge. 1992). In general, estuaries tend to be N limited [(Paerl
and Piehler. 2008; Elser et al.. 2007; Howarth and Marino. 2006; NRC. 2000; Nixon.
1995; Howarth. 1988; D'Elia et al.. 1986) Appendix 10.1.31; however, some estuaries are
P limited, or colimited by N and P, or switch seasonally between N and P (Howarth et al..
2011; Paerl and Piehler. 2008; Howarth and Marino. 2006).
Altered biogeochemical processes associated with N loading (Appendix 7) may affect
aquatic biota in a diversity of coastal habitats. Habitats associated with coastal areas
include shallow open waters, sandy beaches, mud and sand flats, rocky shores, oyster
beds, coral reefs, mangrove forests, river deltas, tidal pools, and sea grasses (U.S. EPA.
2016g). Appendix 11 will cover wetland ecosystems, including those located on coasts in
which soils and/or sediments are periodically inundated by tides or flooding. Estuaries in
the U.S. are located on both coasts with varying levels of eutrophication occurring
lYBricker et al.. 2007):Figurc 10-21. SAV, including the eelgrass Zostera marina, are
important ecological communities found within some coastal bays and estuaries that are
sensitive to elevated nutrient loading (Appendix 10.2.5). Estuaries provide breeding
grounds, nurseries, and shelter for aquatic biota. Near-coastal coral reefs have a more
limited distribution in the U.S. occurring off south Florida, Texas, Hawaii, and U.S.
territories in the Caribbean and Pacific. Elevated N loading appears to play a role in
susceptibility of coral species to disease and bleaching (Appendix 10.4.2).
In the 2008 ISA and the National Estuarine Eutrophication Assessment (NEEA), a
comprehensive survey of eutrophic conditions in the Nation's estuaries conducted by the
National Oceanic and Atmospheric Administration (NOAA), the most eutrophic estuaries
in the U.S. occurred in the mid-Atlantic region, and the estuaries with the lowest degree
of eutrophication were in the North Atlantic lYBricker et al.. 2007) and Figure 10-21. In
the NEEA, 65% of assessed estuaries had moderate to high overall eutrophic conditions
(Bricker et al.. 2007). In the U.S., Chesapeake Bay is perhaps the best-documented case
study of the effects of human activities on estuarine eutrophication. Human disturbances,
such as landscape changes, have exacerbated the negative impacts of N deposition by
reducing N removal and retention in the upper watershed region. Anthropogenic N inputs
have substantially altered the trophic condition of Chesapeake Bay over the last 50 to
100 years. Signs of eutrophication in the bay include high algal production, low
biodiversity, and large hypoxic and anoxic zones. Other estuaries identified in the 2008
10-4
DRAFT: Do Not Cite or Quote

-------
1	ISA where the extent of hypoxia and algal blooms have increased included Long Island
2	Sound, the Pamlico Estuary in North Carolina, and along the continental shelf adjacent to
3	the Mississippi and the Atchafalaya River discharges to the Gulf of Mexico (U.S. EPA.
4	2008a). This appendix updates the state of the science on N-nutrient enrichment, focusing
5	on U.S. waters, since the release of the 2008 ISA.

1
i
8
0
1
¦
J
o

0


2 40


~ 30
in
O)
o 20
k.

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
the atmosphere. In estuaries adjacent to areas of coastal upwelling, such as some
locations along the Pacific coast of the U.S., oceanic inputs of nutrients may also
represent an important source of N (Brown and Ozretich. 2009). N inputs can be
attributed to point sources from a single outfall or discharge point and nonpoint sources,
which are diffuse (U.S. EPA. 2008a). Both point and nonpoint sources have been
identified as targets for control of N inputs in coastal systems (Stephenson et al.. 2010;
Paerl et al.. 2002).
10.1.2. Trends in Atmospheric Deposition of Nitrogen (N)
Trends in atmospheric deposition are discussed in Appendix 2. In summary, in many
parts of the U.S., including the Southeast, Mid-Atlantic, and Midwest, deposition of
reduced N has increased relative to oxidized N in last few decades, shifting from a nitrate
(NO, ) dominated to an ammonium (NH4) dominated condition. This trend is expected
to continue in the future under existing emission controls and current projections of
atmospheric deposition (Li et al.. 2016d; Ellis et al.. 2013; Pinder et al.. 2008; U.S. EPA.
2008a). Wet deposition is now primarily NH4+ at nearly 70% of U.S. air monitoring
locations and reduced N dominates dry deposition in most parts of the country (Li et al..
2016d). This relative increase in NH4 is attributed to intensified industrial-scale animal
operations, increased application of fertilizer, and successful NOx emission controls (Li
et al.. 2016d; Xing et al.. 2013). Mobile source emissions, especially emissions from
diesel vehicles, which increasingly use urea to control NOx emissions, also contribute to
total reduced N loading as well as near-road runoff in regions receiving heavy traffic
(Bettez et al.. 2013). The increase in highly bioreactive reduced N from deposition and
other sources is often a preferred form of N for phytoplankton, including harmful species
(Appendix 10.3.3). In coastal areas of the U.S., atmospheric inputs are heterogeneous
ranging from <10 to approximately 70% of the N inputs (Table 7-8).
10.1.3. Nitrogen Limitation
As reported for several decades prior to the 2008 ISA, N enrichment of marine and
estuarine waters can alter the ratios among nutrients such as P and Si and affect overall
nutrient limitation. N is the most common limiting nutrient in estuarine and coastal
waters and continued inputs have resulted in N over enrichment and subsequent
alterations to the nutrient balance in these systems (U.S. EPA. 2008a; Elser et al.. 2007;
Howarth and Marino. 2006; NRC. 2000; Paerl et al.. 2000; Nixon. 1995; Howarth. 1988;
D'Eliaetal.. 1986). As described in the 2008 ISA, nutrient limitation may shift along the
estuarine to ocean continuum (U.S. EPA. 2008a). For example, Chesapeake Bay exhibits
June 2018
10-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
a shift in nutrient limitation with P most commonly limited at upstream freshwater
locations. At the transition between fresh and salt water, N and P may be colimiting,
whereas the saltwater environments in the lower bay and sound regions are usually N
limited (Fisher et al.. 1998; Rudek et al.. 1991). Levels of N limitations are affected by
seasonal patterns in estuaries, with N limited conditions likely occurring during the peak
of annual productivity in the summer (U.S. EPA. 2008a').
Since the 2008 ISA, N limitation in estuarine and marine systems has been further
evaluated, with recognition of the shifting nature of nutrient limitation based on relative
nutrient inputs and other ecosystem conditions. The rate of nutrient delivery, especially
N, to coastal waters is strongly correlated to primary production and phytoplankton
biomass (Paerl and Piehler. 2008). Strong input controls causing low riverine P inputs
have, in some cases, exacerbated N limitation downstream, with implications for
estuarine eutrophication (Paerl. 2009). Likewise, if N inputs are high enough, some
estuaries and coastal marine ecosystems can become P or Si limited (Howarth et al..
2011; Paerl and Justic. 2011; Paerl and Piehler. 2008). Under this scenario, if a system
becomes P limited, the N input may travel farther away from its sources and contribute to
eutrophication at greater downstream distances (Howarth et al.. 2011).
The role of N inputs from upstream and the connectivity between freshwater and
receiving estuarine and coastal waters have led to recommendations to reduce both N and
P in upstream waters (Glibert and Burford. 2017; Paerl et al.. 2016b; Woodland et al..
2015; Paerl et al.. 2014; Conlev et al.. 2009; Paerl. 2009). Increased N inputs may be
affecting N limitation in the open ocean as well as in near-coastal areas. (Kim et al..
2014a) reported a detectable increase in NOs concentration in the upper Pacific Ocean.
The rate of increased N relative to P was highest near the source of anthropogenic
emissions in northeastern Asia with rates decreasing eastward across the upper North
Pacific Ocean. The authors suggest that increased N deposition may enhance primary
production and potentially lead to a shift from N to P limitation in this region.
Additional studies support observations reported in the 2008 ISA of N and P dynamics in
estuaries. P limitation can play an important role, particularly seasonally if the water
body has a significant freshwater source (Kemp et al.. 2005; Malone et al.. 1996). P
limited conditions often exist in the low salinity regions while N limitation is more
common downstream in higher salinity waters (Paerl and Otten. 2013). Studies in the
Chesapeake Bay and in Europe have shown that the phytoplankton community in coastal
ecosystems can be P limited over several months in the spring and switch to N limitation
for periods of time as short as 1 week later in the season (Trommer et al.. 2013; Malone
et al.. 1996). In the low-nutrient waters of North Carolina's Alligator River estuary,
phytoplankton primary productivity and biomass (chlorophyll a) increased with N
10-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
additions, but both indicators showed the greatest increases after treatment with
combined N and P, indicating that this system was colimited by N and P (Rossignol et al..
2011). N additions in this region are known to lead to poor water quality and related
complications, such as algal blooms, hypoxia/anoxia (water with DO that is too low to
support marine biota), fish kills, and to impact ecosystem services, including fisheries
and recreation (Pacrl and Piehler. 2008).
10.1.4. Characteristics of Coastal Systems Sensitive to Eutrophication
Each coastal system has site-specific characteristics that influence ecological response to
nutrient loading. A variety of factors that govern the sensitivity of estuaries and
near-coastal marine waters to eutrophication from atmospheric N deposition are
summarized in the 2008 ISA. Of critical importance is the total N input from all sources,
including both atmospheric and nonatmospheric sources (Appendix 10.1.1). Other key
elements include the flushing rate and dilution capacity of the watershed which reflects
the volume of water available to dilute added N (NRC. 2000; Brickeretal.. 1999). The
NEEA defined susceptibility as an estimate of the natural tendency of an estuary to retain
or flush nutrients (Bricker et al.. 2007). In estuaries that have longer residence times,
nutrients are more likely to be taken up by algae and lead to eutrophic conditions.
As described in the 2008 ISA, the principal watershed features that control the amount of
increased N flux to estuaries in the U.S. include human population, agricultural
production, and the size of the estuary relative to its drainage basin (Fisher et al.. 2006;
Caddy. 1993; Peierls et al.. 1991). A study included in the 2008 ISA reported a strong
correlation between population density (persons/km2) and the total N loading from
watershed to estuary (r2 = 0.78) for coastal watersheds in the U.S. (Turner et al.. 2001).
This finding is likely due to the prevalence of automobiles in heavily populated areas,
along with their associated N emissions and deposition, plus the myriad of
nonatmospheric sources of N from human activities, particularly wastewater discharges.
The study authors also determined that direct atmospheric deposition becomes
increasingly more important as a contributor to the total N loading to an estuary as the
water surface area increases relative to total watershed area (terrestrial plus water
surfaces). Because of human population growth and the preference for living in coastal
areas, there is substantial potential for increased N loading to coastal ecosystems from
both atmospheric and nonatmospheric sources.
Freshwater inputs to coastal areas and subsequent response to nutrient inputs depend on
residence time of the nutrient-laden freshwater and the degree of tidal exchange within
the estuary (Zaldivar et al.. 2008). Residence time is defined by physical and hydrological
June 2018
10-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
characteristics of the watershed, such as surface area, volume, depth, and flushing rate
(Pacrl et al.. 2002). In water bodies with short residence times, there is little opportunity
for nutrients to be taken up and for algal blooms to develop (Bricker et al.. 2007). For
example, in the heavily N loaded lower Hudson River estuary, phytoplankton are flushed
away as fast as they can grow due to high input of freshwater and high rates of flushing
(Howarth and Marino. 2006; Howarth et al.. 2000). In the NEEA, systems with longer
flushing times were considered more susceptible to eutrophication (Brickcr et al.. 2007).
Other factors within the highly variable estuarine environment (Appendix 7.2.2) that
influence the composition of biological communities include salinity, DO, and suspended
solids, which vary spatially and temporally along the estuary continuum (Borja et al..
2012). Mixing depth, temperature, and light penetration depth also affect biological
response (Pacrl et al.. 2002). A stratified water column that separates well-oxygenated
surface water from bottom water and sediments is generally required to form hypoxic
zones (Jcwctt et al.. 2010). In contrast to the seasonal or persistent nature of hypoxic
zones, diel-cycling hypoxia typically occurs over hours to days and does not require
stratification. Precipitation events like storms and floods or drought can also modulate
nutrient effects I (Pacrl and Piehler. 2008); Appendix 131.
In the 2008 ISA and NEEA, estuaries characterized as eutrophic were generally those that
had large watershed-to-estuarine surface area, high human population density, high
rainfall and runoff, low dilution, and low flushing rates (U.S. EPA. 2008a; Brickcr et al..
2007). In literature reviewed for this ISA, many studies have re-emphasized the important
role that physical and hydrologic factors play in determining which estuaries and coastal
ecosystems are the most sensitive to N enrichment (Hart et al.. 2015; Wilkerson et al..
2015; Glibert et al.. 2014; Rothenberger et al.. 2014; Howarth et al.. 2011; Kennison et
al.. 2011). The hydrodynamics of a system may play an overriding role in controlling
phytoplankton growth (Hart et al.. 2015; Yang et al.. 2008). These factors, which affect
estuarine response to nutrient loading, should be taken into account when establishing N
input thresholds so that eutrophication can be controlled for different ecosystem types,
hydrologic conditions, and future climate scenarios (Paerl et al.. 2014).
Bricker et al. (2014) evaluated N inputs to the Potomac River Estuary and found that
despite some improvements in the upper estuary (i.e., increased DO and decreased
chlorophyll a in the tidal fresh zone; continued regrowth of sea grasses), eutrophic
conditions have worsened in the lower estuary since the early 1990s. Eutrophic
conditions in the Potomac were found to be representative of the Chesapeake Bay region
and other U.S. estuaries, with moderate to high levels of nutrient-related degradation,
particularly compared with similar river-dominated low-flow systems (Bricker et al..
2014). These river-dominated watersheds with low flow (>10 days residence time), high
10-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
population density (i.e., >100 people/km2), and >40% of land use classified as urban
and/or agriculture were predicted to be likely impacted by eutrophication.
(Scavia and Liu. 2009) evaluated a nutrient-driven phytoplankton model that provides a
first-order screening tool for estuarine susceptibility classification. Using data from
75 estuaries, they found that the susceptibility of an estuary to nutrient loading could be
estimated based on the ratio of river inflow (Q) to estuarine volume (V). In this analysis,
efficiency appeared to decrease roughly with the inverse square root of OIV\
s = 0.908(6>/F) "47(/^2 = 0.53), where s represents mean values arising from the
75 estimated normal distributions. Model results showed that estuaries with a 0: F value
greater than 2.0/year are less susceptible to nutrient loads, and those with Q: V values
between 0.3 and 2.0/year are moderately susceptible. Case studies showed that Q:V—and
thus estuarine sensitivity to nutrient loading—can vary between seasons and with storm
events due in part to fluctuations in river inflow.
In some estuaries, especially in the Pacific Northwest, upwelling and oceanic exchange
caused by regional wind patterns likely control primary production rather than
anthropogenic nutrient loading (Brown and Ozretich. 2009; Hickev and Banas. 2003).
Nutrient inputs from local and regional upwelling in these systems can be difficult to
discern from anthropogenic sources. Upwelling-dominated areas are characterized by
short water residence time and have a moderately low expression of eutrophication
symptoms, although nutrient concentrations are high (Kaldv et al.. 2017; Brown and
Ozretich. 2009; Bricker et al.. 2008; Bricker et al.. 2007). Transfer of hypoxic water from
upwelling to estuaries can also occur but is not linked to anthropogenic nutrient additions
(Brown and Power. 2011).
10.1.4.1. Climate Modification of Ecosystem Response to Nitrogen
Climate-related changes including temperature, precipitation, wind patterns, extreme
weather events, stronger estuary stratification, increased metabolism and organic
production, and sea level rise are all expected to modify coastal habitats (Altieri and
Gedan. 2015; Statham. 2012; Rabalais et al.. 2009). These interacting factors will alter
sensitivity to N loading and ecosystem response to nutrient inputs. For example,
eutrophic conditions and the extent and duration of hypoxia are predicted to increase with
changes in temperature and precipitation (Altieri and Gedan. 2015; Rabalais et al.. 2009;
Boesch et al.. 2007). Freshwater and nutrient contributions to estuaries are expected to
rise due to predicted increases in surface water flow and runoff from watersheds
(Rabalais et al.. 2010; Adrian et al.. 2009; Whitehead et al.. 2009). Howarth et al. (2012)
demonstrated larger N fluxes (larger percent delivery of human N inputs) in wetter
June 2018
10-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
climates with more discharge, across 154 different watersheds in the U.S. and Europe.
High organic loads and freshwater inputs associated with extreme weather events may
enhance stratification and contribute to hypoxia ("Wetz and Yoskowitz. 2013).
Temperature modification leading to sea level rise and inputs of freshwater will likely
alter salinity gradients and increase stratification within estuaries (Statham. 2012; Najjar
etal.. 2010V
In estuaries affected by nutrient enrichment and climate-associated alterations described
above, conditions favor a shift in phytoplankton toward greater abundance and
distribution of toxic cyanobacteria associated with increased prevalence of HABs and
declines in SAV (Pacrl et al.. 2016a; Najjar et al.. 2010). Research indicates thatN and
climate change will interact to drive losses in biodiversity that will be more than additive
compared to each independent force (Porter etal.. 2013). Decreases in pH associated with
nutrient-enhanced coastal eutrophication combined with elevated atmospheric CO2 could
increase susceptibility of fauna to ocean acidification I (Cai et al.. 2011c).
Appendix 10.51. Coral reef ecosystems are particularly susceptible to combined effects of
acidification, rising sea level, warming trends, and eutrophication. Appendix 13 includes
a more detailed discussion of how climate (e.g., temperature and precipitation) modifies
ecosystem response to N loading.
10.2. Indicators of Nutrient Enrichment
In the complex environment of the freshwater-to-ocean continuum there are many
chemical and biological indicators of eutrophic conditions (Table 7-8). One approach is
to measure total nutrient loading and concentrations; however, these data need to be
interpreted in the context of the physical and hydrological characteristics that determine
ecosystem response. Water quality measures (water clarity, DO) and quantification of
nutrients along with biological indicators such as chlorophyll a, phytoplankton
abundance, HABs, macroalgal abundance, SAV, and fish kills can all be used to assess
responses to nutrient loading. Some biological indicators such as chlorophyll a are
directly linked to nutrient enrichment and provide evidence of early response to added
nutrients, while other indicators such as low DO and decreases in SAV indicate that the
degree of eutrophication has progressed (Borja et al.. 2012). A summary of key indicators
of estuarine eutrophication is provided in Table 10-1. and these indicators are discussed
in the following sections. The 2008 ISA used the five ecological indicators shown in
Figure 10-3 (chlorophyll a, harmful/nuisance/toxic algal blooms, macroalgae, DO, and
SAV) included in the Assessment of Estuarine Trophic Status (ASSETS) categorical
Eutrophication Condition Index (ECI) in the NEEA to estimate the likelihood that an
June 2018
10-11
DRAFT: Do Not Cite or Quote

-------
estuary is experiencing eutrophication or will experience eutrophication (Brickcr et al..
2007V
Table 10-1 Indicators of Estuarine Eutrophication.
Indicator	Description
Chlorophyll a	Excess N input will stimulate primary productivity, and chlorophyll a concentration is an
indicator of phytoplankton biomass. Phytoplankton biomass is strongly controlled in
coastal waters by the availability and supply rates of nutrients, especially N (Paerl and
Piehler. 2008).
Excess N input can cause nuisance or toxic algal blooms, which release toxins in the
water that can poison aquatic animals and threaten human health (Baron et al.. 2012).
Problem occurrences of HABs increased 13% from 1997 to 2007 in Mid-Atlantic
estuaries, the only region for which data were available, along with increasing N loading
(Bricker et al.. 2008).
Macroalgal abundance Macroalgal blooms can cause the loss of important submerged aquatic vegetation by
blocking sunlight, and opportunistic macroalgae can outcompete many seagrass
species (Kennison et al.. 2011). They can also cause hypoxia and smother seagrass,
coral, clams/oysters, and other benthic organisms. High biomasses also cause "stinky
beach" by rafting onto beaches and decaying.
Dissolved oxygen (DO) Dissolved O2 concentration decreases with increasing algal abundance under elevated
N because microbes consume Chas they decompose dead algae. Oxygen depletion
mainly occurs in bottom waters under stratified conditions. Increased atmospheric N
deposition combined with N loading from other sources will likely affect the size,
frequency, and severity of hypoxia events (Rabalais et al.. 2010). Hypoxia also
contributes to ocean acidification, which is detrimental to the growth and development
of calcifying organisms (Howarth et al.. 2011). The largest documented zone of hypoxic
coastal water in the U.S. is located in the northern Gulf of Mexico (Dale et al.. 2010).
Excess N input stimulates algal growth, and opportunistic macroalgae may block the
penetration of sunlight into the water column and outcompete seagrasses, leading to
reduced SAV coverage. Increased epiphyte loads on the surface of macrophytes from
nutrient enrichment may reduce biomass, shoot density, percentage cover, production
and growth of SAV (G.Nelson. 2017). Reduced extent of eelgrass was found to
correspond to increased loading of N (from wastewater, fertilizer, and atmospheric
deposition) to small to medium shallow estuaries in New England (Latimer and Reqo,
2010), and the distribution of SAV in Chesapeake Bay is used as an indicator of
diversity and biological balance in the U.S. EPA Report on the Environment (U.S. EPA.
2016k).
Chi = chlorophyll; DO = dissolved oxygen; EPA = Environmental Protection Agency; HAB = harmful algal blooms; N = nitrogen;
02 = oxygen; SAV = submerged aquatic vegetation.
Harmful/nuisance/toxic
algal blooms (HABs)
Submerged aquatic
vegetation (SAV)
June 2018
10-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
S.R Brickcr et at./Harmful Algae S (2008) 21-32
Impact: No Problem /low Moderate low
3
UJ
Moderate
Moderate high
High
\.\Y
MM
Few symptoms Symptoms occur Symptoms occur Symptoms occur Symptoms occur
occur 31 more than episodically and/or less regularly less regularly and/or periodically or
minimal levels over a small to and/or over a over a medium to persstendy arxl/or over
medium area medium area extensive area an extensive area
Influencing factors
ds and susceptibility)
Key to symbols
Submerged

aquatic vegetation
Chlorophyll a
Nuisance/toxic
blooms (HAB)
Macroalgae
Dissolved oxygen
Source: Bricker et al. (2008).
Figure 10-3 Biological indicator responses to nutrient enrichment.
10.2.1. Chlorophyll a
Chlorophyll a is part of the light harvesting complex (group of pigments) used by
photosynthetic organisms to convert light energy into carbohydrates through a series of
biochemical reactions. The concentration of Chlorophyll a is often used as a proxy for
phytoplankton biomass. Algae form the base of the coastal food web and excess algal
growth is often directly linked to nutrient enrichment. This indicator can be easily
quantified, and linked to aircraft and satellite-based remote sensing and autonomous
monitoring platforms to assess effects at the regional and ecosystem level
(Appendix 7.2.7). Chlorophyll a is widely used to assess eutrophic conditions because of
its sensitivity to nutrient inputs and was one of the indicators of overall eutrophic
condition of U.S. coastal areas in the 2008 ISA (U.S. EPA, 2008a). Increased levels of
chlorophyll a can signal an early stage of water quality degradation related to nutrient
loading and measurements of chlorophyll a are incorporated into water quality
monitoring programs (Appendix 7.2.7 and Appendix 7.2.10). High concentration of
chlorophyll a suggests that algal biomass is sufficiently high to contribute to low DO
concentration from increased decomposition of dead algae. Due to the strength of
hydrologic forces (i.e., freshwater inputs and tidal flushing) in some types of coastal
systems, benthic chlorophyll a may be a more sensitive indicator of ecosystem response
to N enrichment than planktonic chlorophyll a in well-flushed systems (Paerl and Piehler.
June 2018
10-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
2008). Phytoplankton blooms can also be advected into estuarine systems from the
coastal ocean (Brown and Ozretich. 2009).
Chlorophyll a concentrations are commonly included in standardized frameworks of
eutrophic condition [Appendix 10.2.6; (Borjaetal.. 2012)1. Many of these indices define
spatial extent and temporal sampling periods and statistical measures to determine
representative concentrations. Several chlorophyll a thresholds have been identified for
U.S. water bodies (Table 10-2). The U.S. EPA National Coastal Condition Assessment
(NCCA; formally known as National Coastal Assessment) measures chlorophyll a from
an annual index period (June to October) and compares the samples to reference
conditions to determine a rating [>20 |ig/L. poor; 5-20 |ig/L. fair; <5 |ig/L. good; (U.S.
EPA. 2016g. 2012bVI. For the NEEA ASSETS, the 90th percentile of annual values for
chlorophyll a combined with spatial coverage and frequency of occurrence of blooms are
reported for distinct salinity zones (tidal fresh 0-0.5 ppt), mixing zone (0.5 to 25 ppt),
and seawater (>25 ppt); 0-5 |ig/L is a water body with low risk of eutrophication,
5-20 |ig/L is moderate, >20 |ig/L is high (Bricker et al.. 2007).
Table 10-2 Chlorophyll a thresholds used in methods to evaluate the status of
phytoplankton in U.S. coastal and estuarine water bodies.
Method/
Approach
Chlorophyll a
Thresholds
and Ranges
(ng/L)
Sample
Time
Frame
Statistical
Measure
Other
Characteristics
Community
Composition
Indicators in
Overall
Eutrophication
Index
U.S. EPA
NCCA
Poor >20
Fair 5-20
Good 0-5
Index
period
(June-Oct.)
Concentration
percentage of
coastal area in
poor, fair, and
good condition
based on
probabilistic
sampling design
for 90% conf. in
areal result

No
Chlorophyll a,
water clarity, DO,
DIP, DIN
ASSETS
(eutrophic
condition
component
only)
Poor >20
Fair 5-20
Good 0-5
Annual
90th percentile
Chlorophyll a
concentration of
annual data
Spatial
coverage,
frequency
occurrence
Nuisance and
toxic bloom
occurrence,
frequency,
duration
Chlorophyll a,
macroalgae, DO,
seagrasses,
nuisance/toxic
algal blooms
ASSETS = Assessment of Estuarine Trophic Status; conf. = confidence; DIN = dissolved inorganic nitrogen: DIP = dissolved
inorganic phosphorus; DO = dissolved oxygen; EPA = Environmental Protection Agency; L = liter; NCCA = National Coastal
Condition Assessment.
Source: Boria et al. (2012).
June 2018
10-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
In the NEEA, elevated chlorophyll a concentration was the most widespread documented
symptom of eutrophication (Bricker et al.. 2007). Half of the estuaries for which data
were available exhibited high chlorophyll a concentration (Bricker et al.. 2007). In the
2008 ISA, San Francisco Bay, CA was an example of an estuary that has experienced
considerable increases in chlorophyll a concentrations in recent years.
Phytoplankton biomass, as indicated by chlorophyll a concentration, is strongly
controlled in estuaries by the availability and supply rates of nutrients, especially N (Pacrl
and Piehler. 2008). Bioassays conducted in the low-nutrient Alligator River estuary in
North Carolina showed that N enrichment is directly related to increasing chlorophyll a
concentration. Although the highest increase occurred in response to addition of both N
and P, dissolved inorganic N (DIN) treatment alone stimulated chlorophyll a in some
treatments (Rossignol et al.. 2011). A 3-year data set from Raritan Bay, NJ indicates that
nutrient loading contributed to high concentrations of chlorophyll a from 2010-2012
(Rothcnbcrgcr et al.. 2014). Winter and spring N loading to Mattawoman Creek estuary
and other shallow estuaries in the Chesapeake Bay region was highly correlated to
summer chlorophyll a concentrations (Bovnton et al.. 2014). However, after point source
nutrient (N and P) reductions were enacted, a decrease in chlorophyll a was observed in
Mattawoman Creek. In North Carolina's Neuse River estuary, which is part of the
Albemarle-Pamlico estuary, it appears that the elevated loading of total N (TN)
contributed to higher annual average chlorophyll a values from 2000-2009 (Lcbo et al..
2012).
While recent research indicates that N loading remains a strong predictor of chlorophyll a
concentrations under most conditions, there has been increasing discussion regarding the
role of other factors in altering the strength of this relationship. The impact of nutrient
inputs can at times be overtaken by hydrologic features, seasonal variations, climatic
changes, and oscillations exerting greater control over phytoplankton dynamics (Pacrl et
al.. 2010). In Buzzard's Bay, MA, both nutrient loading and shape of the embayment
were factors in determining spatial and temporal water quality trends including the
observation that chlorophyll a is increasing at a faster rate than N enrichment (Rheubanet
al.. 2016). It appears that more chlorophyll a per unit of TN was produced as the 22-year
time series progressed. In Corpus Christi Bay, TX, rainfall events altered nutrient N:P
ratios but did not affect chlorophyll a (Turner et al.. 2015). Instead, chlorophyll a
concentration in weekly samples mainly followed seasonal trends by increasing in spring
and summer and decreasing in fall and winter. No significant relationships were observed
between annual TN, total phosphorus (TP) load, and annual mean chlorophyll a
concentrations in the Guana Tolomato Matanzas estuary in Florida, rather phytoplankton
biomass in this well-flushed system was influenced by temperature, precipitation, water
residence times, and tidal exchange (Hart et al.. 2015). (Glibert et al.. 2014) found
10-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
significant increases in chlorophyll a over time at only three out of seven study areas in
the Maryland/Virginia coastal lagoon, although regionally chlorophyll a concentrations
increased due to increasing anthropogenic nutrient loads and increased freshwater flow in
the early 2000s. In the Neuse River estuary, NC, which has a long residence time
allowing for detection of nutrient stimulation, nutrient addition bioassays along the
estuary indicated strong N limitation at the chlorophyll a maximum and downstream
where there was a strong preference of NH44" over NOs as a DIN source (Paerl and
Piehler. 2008).
Using satellite and meteorological data from U.S. Atlantic coastal waters, (Kim et al..
2014b) found that precipitation events were associated with increased levels of
chlorophyll a in low nutrient areas (defined as having NO3 concentrations of less than
1 |iM). but precipitation was correlated with lower chlorophyll a concentrations in high
nutrient areas (defined as having NO3 concentrations greater than 1 |iM). This is likely
because in low nutrient areas, new N input from precipitation stimulated phytoplankton
growth. In areas already high in nutrients, wind associated with precipitation events may
have deepened the mixed layer and the resulting loss of light availability caused
chlorophyll a concentrations to decline (Kim et al.. 2014b). Two studies modeling
historical data from Chesapeake Bay indicate that variation in climatic conditions
dominated phytoplankton dynamics in the bay in recent years (Harding et al.. 2016a;
Harding et al.. 2016b). Much of the bay is producing more chlorophyll a per unit TN than
in the past, leading Harding and others to suggest that return to the historical relationship
between N and chlorophyll a is unlikely.
In four coastal estuaries that shifted from eutrophication to oligotrophication (lower
nutrients), the degree of return of chlorophyll a to reference status varied, likely due to
concurrent changes in the estuaries from costressors and the time elapsed from
eutrophication to nutrient reduction (Duarte et al.. 2009). Due to these shifting baselines
the authors suggest that the current paradigm of nutrient reduction to a historical level
needs to be replaced by targets that maintain key ecosystem functions in the context of
changing conditions in the estuaries over time.
10.2.2. Harmful/Nuisance/Toxic Algal Blooms
Nuisance or toxic algal blooms reflect the proliferation of a toxic or nuisance algal
species that negatively affects natural resources or humans. Blooms are increasing in
outbreak frequency and extent in the U.S. and other countries, and N is one of the
nutrients known to promote HAB formation (Baron et al.. 2012; Heisler et al.. 2008). A
concurrent increase in atmospheric sources of N to coastal areas and proliferation of
June 2018
10-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
harmful algal bloom formation have been recognized for several decades (Paerl et al..
2002; Paerl and Whitall. 1999; Paerl. 1997). The form of N delivered to coastal regions
of the U.S. from atmospheric and other sources is changing from primarily NOs to an
increase in reduced forms of N (Appendix 10.1.2). which are favored by some HAB
forming species (Glibert et al.. 2016). HABs caused by N enrichment can release toxins
that are harmful to fish and shellfish and that may accumulate in predators and organisms
higher in the food web (Johnson et al.. 2010). For example, cyanobacteria blooms
(cyanoHAB) have been documented in estuaries along the U.S. Mid-Atlantic and
Southeast coasts as well as Gulf coast estuaries (Preece et al.. 2017). Other blooms are
not toxic but may cause low DO events due to very high biomass. In addition to effects
on biota, HABs cause a range of responses in humans from direct dermatitis, such as
swimmers itch, to severe food poisoning, liver and kidney toxicity, and paralysis (Peel et
al.. 2013; Johnson et al.. 2010).
The frequency and duration of algal blooms is one of five indicators used in the NEEA
and was included as an indicator of eutrophic conditions in the 2008 ISA (U.S. EPA.
2008a; Bricker et al.. 2007). Of the 81 estuary systems for which data were available in
the NEEA, 26 exhibited a moderate or high symptom expression for nuisance or toxic
algae (Bricker et al.. 2007). In the previous 2008 ISA review, the frequency of
phytoplankton blooms and the extent and severity of hypoxia were documented to have
increased in the Chesapeake Bay (Officer et al.. 1984) and Pamlico Sound estuaries in
North Carolina (Paerl and Piehler. 2008) and along the continental shelf adjacent to the
Mississippi and Atchafalaya river discharges to the Gulf of Mexico (Eadie et al.. 1994).
New tools and monitoring approaches to further characterize HABs have become
available since the 2008 ISA (Appendix 7.2.7). For example, solid phase adsorption toxin
tracking (SPATT), a passive sampling tool which simulates the contamination of
filter-feeding bivalves, has been used recently in the field to detect the presence of algal
toxins with greater accuracy than grab-sampling methods (Gibble and Kudela. 2014;
Kudela. 2011; Lane et al.. 2010). Remote sensing systems are increasingly being used to
forecast and monitor HABs in coastal waters (Klemas. 2012).
Harmful effects of HAB toxins on fish and wildlife are readily transferred through the
food web and may persist even when the bloom conditions have passed. Microcystin, a
class of toxins produced by many cyanoHABs is found in all trophic levels from shellfish
to finfish to top level predators and can persist in the foodweb for months (Preece et al..
2017).(Wood et al.. 2014) found that microcystin persisted in overwintering populations
of estuarine finfish, common wedge clam (Rangia cuneata) and blue crab (Callinectes
sapidus) in the James River estuary, VA although the highest tissue concentrations and
greatest percentage of individuals affected were observed when toxins in the water
column were at maximum levels. The toxin was present in both muscle and viscera of
10-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
blue crabs at concentrations that have been shown to have physiological effects on other
species of estuarine crab (Wood et al.. 2014). Accumulation of microcystin measured
over the course of 2 years in the James River estuary, near Chesapeake Bay, was found to
be highest in suspension feeding animals while top predators (piscivores), scavengers,
and benthic feeders all had lower levels of microcystin (Bukavcckas et al.. 2017). In
Monterey Bay, CA, deaths of 21 sea otters (Enhydra lutris), a federally listed threatened
species, were attributed to hepatotoxic shellfish poisoning due to trophic transfer of
microcystin observed in this study to have originated from nutrient-impaired lakes and
rivers discharging to the bay (Miller et al.. 2010). Other HAB toxins, such as domoic
acid, are also known to be transferred up the food web (Trainer et al.. 2012). Domoic acid
from recent blooms of the diatom Pseudo-nitzschia along the U.S. West Coast has
impacted razor clam (Siliqua patula) and Dungeness crab (Metacarcinus magister)
fisheries, led to symptoms of domoic acid poisoning in California sea lions (Zalophus
californianus), and was detected in additional marine mammals from southern California
to northern Washington (Mccabe et al.. 2016).
Research on HAB-forming species have shown that the form of N supplied affects
phytoplankton growth (Glibert et al.. 2016). Generally, NH4+ is considered to be the
preferred form of N for some phytoplankton due to lower energy requirements for uptake
and assimilation; however, diatoms specialize in use of oxidized N forms (Glibert et al..
2016). For example, the HAB-forming dinoflagellate species Heterosigma akashiwo is
able to grow well with a pulsed supply of NH/, NOs . and urea under low light
conditions and assimilate stores ofN nutrients, suggesting that this species could
outcompete diatoms in silicate-limited, N enriched coastal areas (Kok et al.. 2015). In
Pseudo-nitzschia spp., laboratory experiments showed that both of the species studied
were able to grow on NH4+, NO3 . and urea, although urea was the preferred form of N
(Melliti Ben Garali et al.. 2016). These findings were supported by a field experiment in
which chlorophyll a concentration significantly increased and exponential growth
occurred in all N-enriched in situ microcosms until the end of the experiment, with
specific growth rates highest in the urea and NO3 additions. The dinoflagellate Akashiwo
sanguinea showed different growth profiles and N assimilation with form of N and
concentration, growing faster in NH4 and with greater enzyme affinity for urea (Liu et
al.. 2015). Differential growth responses of HAB species to reduced N can alter
phytoplankton community composition and biodiversity (Appendix 10.3.3).
There is also evidence, mostly from freshwater systems, that the form of N affects toxin
production of some HAB species (Appendix 9.2.6.1). In nutrient amendment experiments
with water collected from the tidally influenced Transquaking River, which flows into
Chesapeake Bay, Microcystis was stimulated by N more frequently than P, and
abundances of toxic and nontoxic strains were enhanced to different degrees by inorganic
10-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
and organic N (Davis et al.. 2010). Response of Alexandrium fundyense to nutrient
addition varied throughout the course of bloom events in Northport-Huntington Bay, NY
(Hattcnrath et al.. 2010). Addition of NH44" to bloom water most frequently resulted in
statistically significant increases of A. fundyense density and toxin concentration
compared to other forms of N (glutamine, NOs . and/or urea). Davidson et al. (2012)
found that, in general, laboratory studies show that toxin production can be influenced by
nutrient ratios, but extrapolation of those results to the specific species and conditions in
the field is difficult.
Modeling studies have reported on potentially altered future scenarios of phytoplankton
community changes and HAB formation, intensity, duration, and toxicity due to changes
in N deposition (Lee and Yoo. 2016; Glibert et al.. 2010b). Pinder et al. (2008) reported
on modeled scenarios that predict alterations in the frequency, intensity, toxicity, and
species composition of algal blooms due to higher rates of N deposition and changes in
the reduced N to oxidized N ratio. Future changes in HAB dynamics will be affected by
climate change (Appendix 10.1.4.1) and increased N loading, and integrated ecosystem
models that couple the atmosphere, land, and coastal ocean are needed to estimate these
HAB responses (Glibert et al.. 2010a). Long-term monitoring (1987-2005) of
phytoplankton populations in the Bay of Fundy in southwestern New Brunswick, Canada
did not link HABs to nutrients, rather many species abundances are explained by climate
and weather patterns (Martin et al.. 2009). Vox A. fundyense there was a negative
relationship with cell density and NO3 .
Table 10-3 summarizes new studies from U.S. waters on levels and forms of N at which
effects are manifested in phytoplankton.
10-19
DRAFT: Do Not Cite or Quote

-------
Table 10-3 Levels and forms of nitrogen at which effects on phytoplankton are manifest in U.S. coastal waters
evaluated in the literature since the 2008 Integrated Science Assessment for Oxides of Nitrogen and
Sulfur-Ecological Criteria.
Study Site
Ambient N
Deposition
N Additions
Ecological/Biological Effects
Species
Reference
Northport-
Huntington Bay
complex, NY
Addition of NH4+
(10 |jM-40 |jM)
Addition of NhV significantly increased A. fundyense Phytoplankton
densities compared to the control. The addition of NHV (Alexandrium
(40 |jM) yielded a significant increase in both A.	fundyense)
fundyense densities and toxin concentrations four and
8x, respectively, compared to controls.
Hattenrath et al.
(2010)
Raritan Bay, NJ
Ambient levels Multivariate analyses of a 3-yr data set indicated that the Phytoplankton
abundance of HAB species Heterosigma akashiwo is	(Heterosigma
positively associated with NO3" in Raritan Bay. Both	akashiwo and
climatic conditions and nutrient concentrations affect	13 other HABs
phytoplankton bloom composition in the bay.	identified)
Rothenberqer et
al. (2014)
Raritan Bay, NJ
Nutrients (N and
Fe) added alone or
in combination to
different treatments.
N was added as a
single pulse sodium
nitrate (NaNOs) to
increase NO3"
concentrations by
approximately
10 |jM. The N
additions lowered
the Si:N ratios from
~3 to <1
Dinoflagellates and HAB-forming taxa increased to a
greater extent when NO3" levels were high (which led to
a low Si:N ratio of less than one). Centric and
chain-forming diatoms resulted from enriched NO3"
concentrations, differing from the pennate diatoms and
green flagellates that accompanied ambient NO3"
concentrations. Dinoflagellates in the genus Dinophysis,
which could be HAB-forming, also resulted from enriched
NO3" and lowered Si:N.
Phytoplankton
Dinoflagellates,
diatoms
Rothenberqer
and Calomeni
(2016)
June 2018
10-20
DRAFT: Do Not Cite or Quote

-------
Table 10-3 (Continued): Levels and forms of nitrogen at which effects on phytoplankton are manifest in U.S.
coastal waters evaluated in the literature since the 2008 Integrated Science Assessment
for Oxides of Nitrogen and Sulfur-Ecological Criteria.

Ambient N




Study Site
Deposition
N Additions
Ecological/Biological Effects
Species
Reference
Ocean surface
Based on direct

Precipitation events in coastal waters of the eastern U.S.
Phytoplankton
Kim et al.
from 28°N to
measurements of

increased the chlorophyll a concentration up to 15% in

(2014b)
44°N and from
wet deposition along

low-nutrient areas (<1 |jM NO3") but decreased the


the East Coast
the East Coast of

chlorophyll a concentration in nutrient-replete areas


of the U.S. to
the U.S., the N

(>1 |jM NO3"). The authors suggested that in


60-70°W
supply through wet
deposition was
estimated to be
25-45 mmol N/m2/yr

nutrient-depleted areas (south of 36°N), the added
nutrients were a dominant factor increasing the
chlorophyll a concentration, whereas in the
nutrient-replete areas (north of 36°N), where
phytoplankton growth was light limited, reduced light
availability was the dominant factor determining reduced
chlorophyll a concentration.


Matta woman
Used atmospheric

Strong relationships were found between N loading and
Phytoplankton and
Bovnton et al.
Creek,
deposition data from

algal biomass and between algal biomass and water
SAV
(2014)
Chesapeake
Bovnton et al.

clarity. Winter-spring N loading and summer chlorophyll


Bay
(2008). includina N
from both wet and
dry deposition
(0.81 mg N/L as an
annual average
concentration).
Direct atmospheric
deposition to surface
waters of the creek
contributed about
6,000 kg N/yr or
about 16 kg N/day to
the creek system.

a were found to be highly correlated, a relationship which
appears to be linear.


June 2018
10-21
DRAFT: Do Not Cite or Quote

-------
Table 10-3 (Continued): Levels and forms of nitrogen at which effects on phytoplankton are manifest in U.S.
coastal waters evaluated in the literature since the 2008 Integrated Science Assessment
for Oxides of Nitrogen and Sulfur-Ecological Criteria.
Study Site
Ambient N
Deposition
N Additions
Ecological/Biological Effects
Species
Reference
Tidally
influenced
Transquaking
River which
flows into
Chesapeake
Bay
20 |jM (NOs"),
20 |jM NH4+ (NH4+),
10 |jM (= 20 |jM N)
urea, 10 pM
(= 20 |jM N),
L-glutamine (GA), P
(1.25 |jM
orthophosphate), or
a combined
treatment of NO3"
and P
Microcystis was simulated by N more frequently than P,
and abundances of toxic and nontoxic strains were
enhanced to different degrees by inorganic N and
organic N. Toxic Microcystis abundance increased more
with inorganic N than organic N.
Toxic and nontoxic
strains of Microcystis
Davis et al.
(2010)
Maryland and
Virginia coastal
bays
Measurements of
atmospheric
deposition since
2000, based on the
NADP, suggest that
NO3" is decreasing
and NH4+ from
deposition is stable
for the coastal bays
(NADP data
http://nadp.sws.uiuc.
ed u/sites/ntn/NTNtre
nds.html?sitelD=MD
18).
N in the water column is dominated by reduced N
(primarily NH4+) with low concentrations of NO3" also
present, resulting in phytoplankton community shifts to
those species that can do well under such conditions.
Submerged aquatic vegetation has decreased.
Phytoplankton and
SAV
Glibert et al.
(2014)
Alligator River
estuary, NC
Potential for
atmospheric
deposition of N from
nearby farm (wet
NH4+ concentrations
increased greatly
close to farm).
DIN additions:
140 |jg N-NhV/L,
140 |jg N-NO3-/L,
and 70 pg N-NhV/L
plus 70 pg
N-NCV/L
Significant increase in phytoplankton biomass (chl a) and Phytoplankton
rates of primary productivity due to N enrichment. DIN
treatments alone significantly stimulated chl a in two out
of five tests for all three DIN addition treatments,
although DIN + DIP treatments provided the largest
increase in chl a.
Rossiqnol et al.
(2011)
June 2018
10-22
DRAFT: Do Not Cite or Quote

-------
Table 10-3 (Continued): Levels and forms of nitrogen at which effects on phytoplankton are manifest in U.S.
coastal waters evaluated in the literature since the 2008 Integrated Science Assessment
for Oxides of Nitrogen and Sulfur-Ecological Criteria.

Ambient N




Study Site
Deposition
N Additions
Ecological/Biological Effects
Species
Reference
Neuse River

Ambient, acute DIN
Study noted increase in phytoplankton bloom frequency
Phytoplankton
Paerl et al.
estuary, NC

inputs via runoff
and magnitude (indicated by increasing chl a variability)

(2010)


from hydrologic
overtime in response to acute DIN inputs from




pulses (hurricanes,
hydrologic pulses. Control of algal bloom duration,




tropical storms,
thresholds, taxonomic composition, and spatial extent




heavy rainfall
may be dictated by climatic changes and oscillations




events)
instead of nutrient inputs.


New River

N addition to
Varying the form of nutrients promoted growth of
Phytoplankton
Altman and
estuary, NC

estuary water in the
different phytoplankton groups based on photopigment

Paerl (2012)


form of DIN, organic
analysis. Dinoflagellates, chlorophytes, and




N from river water
cyanobacteria responded to dissolved organic N while




(with dissolved
cyanobacteria increases were most frequent with




inorganic P) or urea
inorganic N addition.


Four sites with

NH4+, NOs", and
Along a gradient from highly developed to undeveloped
Phytoplankton
Reed et al.
tidal influence,

urea treatments
sites, phytoplankton communities at the more developed

(2016)
South Carolina

applied separately
sties had higher biomass and growth rate with N


coast

to bioassays. N
(particularly urea) additions and potentially HAB forming




forms were added
species were more often found at the more developed




at Redfield ratios
sites.




(N:P = 16:1)



Ten Mile Creek,

High median
Chl a was negatively correlated with N concentrations
Phytoplankton
Yana et al.
Indian River

concentrations of
and the highest chl a concentrations were related to the

(2008)
Lagoon, FL

total N
conditions of static and open water with long residence




(0.988 mg/L),
time. Hydrodynamics of a system may play an overriding




NO3--N
role in controlling phytoplankton growth.




(0.104 mg/L),





NH4+-N





(0.103 mg/L),





and total Kjeldahl N





(0.829 mg/L)



San Francisco

NOs" and NH4+
A comparison of N isotopes in cells of the HAB species
Phytoplankton
Lehman et al.
Bay/estuary

measured from two
Microcystis aeruginosa with N in rivers flowing into the
(cyanobacteria
(2015)


source estuaries
bay suggested that NH4+, not NCVwas likely the primary
Microcystis




source of N that supported the bloom.
aeruginosa)

June 2018
10-23
DRAFT: Do Not Cite or Quote

-------
Table 10-3 (Continued): Levels and forms of nitrogen at which effects on phytoplankton are manifest in U.S.
coastal waters evaluated in the literature since the 2008 Integrated Science Assessment
for Oxides of Nitrogen and Sulfur-Ecological Criteria.
Ambient N
Study Site	Deposition	N Additions	Ecological/Biological Effects	Species	Reference
San Francisco
Ambient nitrate
The greatest chl a concentration and cell density
Phytoplankton
Lehman et al.
Bay/estuary
ranged from 0.19 to
occurred in the San Joaquin River estuary which had a
(cyanobacteria
(2008)

0.36 mg/L. Ambient
high average nitrate concentration (0.36 mg/L). The
Microcystis


ammonia ranged
second highest chl a concentration and cell density
aeruginosa)


0.02 to 0.06 mg/L.
occurred in the Old River estuary which had relatively
low N concentration (mean 0.19 mg/L). Differences in chl
a were correlated with environmental conditions,
particularly streamflow and water temperature and
secondarily to nutrient concentrations and ratios.


DIN = dissolved inorganic nitrogen; DON = dissolved organic nitrogen; HAB = harmful algal blooms; N = nitrogen; NADP = National Atmospheric Deposition Program;
NH4+ = ammonium ion; N03" = nitrate; P = phosphorus; SAV = submerged aquatic vegetation; W = west.
June 2018
10-24
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
10.2.3. Macroalgal Abundance
The abundance of macroalgae, which are generally referred to collectively as "seaweed,"
was an indicator of eutrophic condition in the 2008 ISA (U.S. EPA. 2008a'). Macroalgae
have been combined with other indicators to classify estuarine condition (Borja ct al..
2012). For example, NEEA includes an assessment of macroalgae (Brickcr et al.. 2007).
Macroalgal blooms can contribute to loss of important SAV by blocking the penetration
of sunlight into the water column. Although macroalgal data for estuaries in the U.S.
were generally sparse, the NEEA reported high macroalgal expression in 15 of the
64 estuaries evaluated (Brickcr et al.. 2007). In some lagoons with limited oceanic
exchange, macroalgae may be a more sensitive biological indicator than phytoplankton
(e.g., McLaughlin et al.. 2014; Nobre et al.. 2005). For example, in most estuaries of the
Southern California Bight, macroalgae is the dominant primary producer and a key
indicator of eutrophication (McLaughlin et al.. 2014). Macroalgae may not be a good
indicator of eutrophication in some upwelling-influenced estuaries in the Pacific
Northwest because an increase in macroalgal biomass in these systems does not appear to
be associated with temporal declines in eelgrass (Hessing-Lewis et al.. 2015; Hessing-
Lewis and Hacker. 2013; Hessing-Lewis et al.. 2011).
Opportunistic, fast-growing macroalgae can exhibit very high rates of N uptake during
periods of high N availability and can often outcompete or block out light for other
macrophytes when N loads are high and variable (Abreu etal.. 2011). This growth of
macroalgae can also smother corals, clams, oysters, and other biota (Bricker et al.. 2007)
and contribute to declines in seagrasses (Olvarnik and Stachowicz. 2012). These
macroalgae also tend to preferentially uptake N in the form of NH4+. In a Danish study,
sea lettuce (Ulva lactuca), which is a common species of macroalgae in U.S. coastal
waters, grew faster and exhibited greater biomass when subjected to NH4 as the N
source compared to NOs (Ale et al.. 2011). This difference was thought to be due to
reduced N being more easily assimilated and used by algae. Similarly, in experimental
manipulations with Gracilaria tenuistipitata, an opportunistic macroalgal species from
China, when both NH4 and NO3 were available, NH4 was assimilated more rapidly and
algal biomass was higher than with NO3 addition alone (Wang et al.. 2014a). Growth of
Caulerpa cylindracea, an invasive macroalga in the Mediterranean Sea, was not inhibited
by high NH3, and was able to outcompete native macroalgae in experimental plots with
nutrient addition (Gennaro et al.. 2015).
Southern California estuaries, some of the most nutrient enriched in the world, have not
been well studied in the past, but recent work is providing important information about
macroalgal growth in response to N loading to these systems. Opportunistic macroalgae
June 2018
10-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
(Ulva intestinalis and Ulva expansa) were shown to take up NOs, from the water very
efficiently at many concentration levels, giving these species the ability to outcompete
other algae in estuaries subject to varying nutrient loads (Kcnnison ct al.. 201 1). Algae
that were depleted in N took up NO3 at higher rates than enriched algae, and uptake rates
slowed as the algae became saturated with N.
However, macroalgal abundance was not found to be directly related to water or sediment
N concentrations in Southern California estuaries (Kcnnison and Fong. 2014). Rather,
macroalgal blooms occurred throughout the year and appeared to be influenced by the
unique physical and hydrological differences between estuaries. For instance, in one
estuary with high nutrient levels in both water and sediment, little to no macroalgal
biomass was present, possibly due to high water velocities that prevented young algal
filaments from attaching(Kcnnison and Fong. 2014). In other estuaries, comparatively
high algal biomass was found year-round even when nutrient supplies were lower than
other sites (although still N enriched), perhaps due to low rates of tidal flushing and
longer water residence times. Internal nutrient processing, variable hydrological regimes,
and multiple external nutrient sources may all contribute to eutrophic conditions
(Kennison and Fong. 2014).
10.2.4. Dissolved Oxygen
DO was included in the 2008 ISA, the NCCA, and the NEEA as an indicator of eutrophic
condition (U.S. EPA. 2008a). Additional information on DO is provided in
Appendix 7.2.3. Oxygen depletion largely occurs only in bottom waters under stratified
conditions, not throughout the entire water column. The decomposition of organic matter
associated with increased algal abundance consumes DO and can reduce DO
concentrations in eutrophic waters to levels that cannot support aquatic life (Jcwctt ct al..
2010). Respiration of microbes, macrophytes, and animal biota can also reduce DO to
very low levels such as primary producers in the dark, deeper portions of the well-mixed
Hudson River estuary (Howarth et al.. 1996b) and as seen every night in a eutrophic
seagrass-dominated system on Cape Cod (Howarth et al.. 2014). In the Cape Cod
seagrass system, hypoxia is common at dawn, following hours of darkness, yet oxygen
levels are supersaturated at the end of the daylight period.
Generally, some biota are impacted at DO levels from 3 to 4 mg/L, and increasingly
adverse effects are observed on biota at lower DO concentrations (Figure 10-4).
Decreased DO can lead to the development of hypoxic or anoxic zones that are
inhospitable to fish and other life forms and can impact ecosystem processes (Diaz et al..
2013; Levin et al.. 2009; Diaz and Rosenberg. 2008). For example, in Chesapeake Bay,
June 2018
10-26
DRAFT: Do Not Cite or Quote

-------
Sturdivant et al. (2014) observed that macrobenthic production was 90% lower during
hypoxic conditions, resulting in abiomass loss similar to 7,320-13,200 metric tons C
over an area of 7,720 km2. The authors estimated this change represented a displacement
of 20 to 35% of macrobenthic activity during the summer.
Mobile Fauna Begin
to Migrate to Higher
DO Areas
100% Saturation
Avoidance by
Fishes
Fishes Absent
Shrimp & Crabs
Absent
Burrowing Stops
Stressed Fauna
Emerge & Lay on
Sediment Surface
Fauna Unable to Escape
Initiate Survival
Behaviors
Mortality of Sensitive
Fauna
Mortality of Tolerant
Fauna
Formation of Microbial
Mats
Sediment Geochemistry
Drastically Altered
No Macrofauna
Survive
Hydrogen Sulfide Builds Up
\ Water Column
DO = dissolved oxygen; I = liter; mg = milligrams.
Source: Diaz et al. (2013V
Figure 10-4 The range of ecological impacts exhibited as dissolved oxygen
levels drop from saturation to anoxia.
June 2018
10-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
In the U.S., the incidence of hypoxia has increased almost 30-fold from 1960 to 2008 and
is now reported in more than 300 systems (Jcwett et al.. 2010). Climate change and
increasing N loading to coastal ecosystems are both expected to widen the distribution
and increase the size of areas affected by eutrophication-induced hypoxia, while also
increasing the frequency and persistence of these events KRabalais et al.. 2010);
Appendix 131. Areas of eutrophication-related hypoxia are found on the U.S. East and
West coasts and the Gulf of Mexico (Figure 10-5). Eutrophication-induced hypoxia,
which has been documented globally, can be characterized by both the duration of the
event and the ecosystem response (Diaz et al.. 2013; Diaz and Rosenberg. 2008).
Summer hypoxia is most common, followed by periodic oxygen (O2) depletion that may
occur in some systems more often than seasonally. In these hypoxic and anoxic areas
("dead zones") only organisms that can live with little or no O2 are present (Diaz et al..
2013; Jewett et al.. 2010). Once an ecosystem reaches severe seasonal hypoxia, there is a
shift to benthic organisms with shorter life spans and smaller body size (Diaz and
Rosenberg. 2008).
Zhou et al. (2014b) estimated the hypoxic volume of Chesapeake Bay over a 25 year
period. The date of onset of hypoxia occurs earlier in the summer and the end of hypoxia
also shifted to earlier in the fall with no trend in the seasonal-maximum hypoxia itself
from 1985 to 2010. Nutrient loading from the Rappahannock, Susquehanna, and Potomac
Rivers explains >85% of the seasonally averaged interannual variability in hypoxic
volumes. Testa and Kemp (2012) investigated interactions between hypoxia and nutrient
cycling in Chesapeake bay, based on analysis of long-term monitoring data collected
during two time periods: 1965 to 1980 and 1985 to 2007. They found that bottom water
in the upper Bay region, where seasonal hypoxia first develops, was enriched in NH4+
and PO4 relative to other regions of the bay. This might help explain the occurrence of
extensive and persistent hypoxia during the month of June even during years of lower N
loading.
10-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
Eutrophic arid Hypoxic Coastal Areas of North America and the Caribbean
Source: http://www.wri.ora/resources/maps/coastal-eutrophic-and-hvpoxic-areas-north-america-and-caribbean modified from Diaz et
al. (2013).
Figure 10-5 Coastal eutrophic and hypoxic areas of North America the
Caribbean.
Recent research has shown that the duration of a hypoxic event is one of the most
important factors affecting macroinvertebrate density and species composition. (Baustian
and Rabalais. 2009) found that hypoxia duration (the number of days during which O2
concentrations were below 1 mg/L) was highly correlated with both reduced species
density and diversity in the northern Gulf of Mexico. Very low O2 levels can be lethal to
fish with no means of escape, and it has been suggested that even sublethal hypoxia may
lower the breeding rate and affect fish populations (Moran et al.. 2010). Hypoxia has
been shown to act as an endocrine disruptor in Atlantic croaker (Micropogonias
imdidatiis) in laboratory studies and biomarkers of reproductive function and endocrine
disruption are observed in field-collected individuals (Murphy et al.. 2009; Thomas and
Rahman. 2009; Thomas et al.. 2007). Hypoxic conditions increase nitric oxide and super
oxide radicals in brain tissue in croakers, causing cellular oxidative damage and inhibited
protein expression in the hypothalamus and leading to neuroendocrine effects (Rahman
and Thomas. 2015). Low O2 conditions have also been shown to alter animal behavior
(Appendix 10.4.1). O2 content influences hatching rate and parental effort among other
reproductive behaviors in three-spined sticklebacks (Gasterosteus aculeatus), so
Eutrophic
#	Hypoxic
#	Sy9tems in Recovery
Dill RwWW Sel*Bn 2D1D.
www oil
June 2018
10-29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
eutrophication-induced hypoxia may alter reproductive output in some fish populations
(Candolin. 2009). However, hypoxia does not appear to negatively affect fisheries below
what would be predicted from N loadings alone, except under circumstances in which
raw sewage is released or when critical habitat is lost for very sensitive species (Breitburg
ct al.. 2009). In a review of long-term chronic effects of hypoxia on commercially
important fishery species (Townhill et al.. 2017) noted that effects range from positive to
negative, and from physiological to ecosystem-level. Some fish can acclimate to hypoxic
conditions and/or take advantage of more susceptible prey, affecting food web dynamics.
Other species cannot avoid hypoxia (especially shellfish) and suffer physiological stress,
mortality, and/or increased predation. Other species have been shown to move away from
the area to avoid stress (Appendix 10.4.1). which also has cascading food web
implications.
The effects of low DO are influenced by the presence of multiple stressors. For example,
(Gobler et al.. 2014) examined concurrent effects of low DO and acidification on the
early lifestages of bay scallops (Argopecten irradians) and hard clams (Mercenaria
mercenaria). Observations in later lifestages of the clams indicated that growth rates
decreased by 40% in combined exposures to hypoxia and acidification. Additional studies
with earlier lifestages indicated effects were more severe with costressors than with either
hypoxia or acidification alone. Juvenile oysters (Crassostrea virginica) grown under
varying conditions of low pH and duration of hypoxia seemed to acclimate and were only
negatively affected by the most severe pH experiment (Keppel et al.. 2016). Growth rates
were reduced 30-37% initially by both brief, repeated hypoxia and long moderate
hypoxia events; however, at the end of the study most oysters were the same size
regardless of treatment. This study also reported that the initial effects on oyster growth
were more pronounced by constant moderate hypoxia (1.3 mg/L) than they were by
severe but cyclical hypoxia (0.5 mg/L).
At the time of the last review, it was documented that the largest zone of hypoxic coastal
water in the U.S., and the second largest in the world, was the northern Gulf of Mexico
Hypoxic Zone on the Louisiana-Texas continental shelf (Dale et al.. 2010; Jewett et al..
2010; U.S. EPA. 2008a). Since the 2008 ISA, observations continue to indicate a large
zone of hypoxia in the northern Gulf of Mexico during the summer months associated
with increased N loading from the Mississippi and Atchafalaya River Basins. The size of
the midsummer bottom-water hypoxia area (<2 mg/L DO) in the northern Gulf of Mexico
has varied considerably since 1985, with a long-term average of 13,751 km2 (5,240 mi2)
(U.S. EPA. 2015c). In the summer of 2017, the hypoxic zone in the Gulf was the largest
ever measured at 14,123 km2 (8,776 mi2) (U.S. EPA. 2017e). Alexander et al. (2008) used
the SPARROW water quality model to show that atmospheric deposition to watersheds in
10-30
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
the Mississippi River Basin is the second largest source of N (16%) to the Gulf, after
effluents from corn and soybean production (52%).
The Hypoxia Task Force (Mississippi River/Gulf of Mexico Watershed Nutrient Task
Force) is a federal/state partnership (five federal agencies, the National Tribal Water
Council and 12 states bordering the Mississippi and Ohio rivers) established in 1997 with
the goal of reducing nutrient inputs and the size of the hypoxic zone in the Gulf of
Mexico (U.S. EPA. 2017d. 2015c). The task force is working toward the goal of reducing
the areal extent of the Gulf of Mexico hypoxic zone to less than 5,000 km2 by 2035, with
an interim target of 20% nutrient load reduction by 2025. This was revised from the 2001
action plan that called for the described reductions by 2015 (U.S. EPA. 2001). Each state
in the task force has a nutrient reduction plan, and progress toward stated goals are
reported in the Mississippi River/Gulf of Mexico Watershed Nutrient Task Force Report
to Congress (U.S. EPA. 2015c) as required by the Harmful Algal Bloom and Hypoxia
Research and Control Amendments Act of 2014 (HABHRC Act. 2014). Other states not
included in the Mississippi River/Gulf of Mexico Watershed Nutrient Task Force have
nutrient reduction programs in place.
Hypoxic conditions are reported from other U.S. locations (Figure 10-5). For example, in
the shallow estuary of Long Island Sound, atmospheric deposition is considered to
comprise a significant fraction of total N loading [19%; (Whitall et al.. 2004)1. and DO
levels below 3 mg/L are common, with levels below 2 mg/L known to occur. During
some years, portions of the Long Island Sound bottom waters become anoxic (DO
<1 mg/L), (Latimer et al.. 2014). The maximum extent and duration of hypoxic events
(<2 mg/L DO) in Long Island Sound has varied considerably since the 1980s (U.S. EPA.
2016k). Between 1987 and 2014, the average annual maximum extent was 98 km2
(61 mi2). In 2014, the hypoxic area was 32 km2 (20 mi2), with the lowest DO levels
occurring in the western end of the Sound. The shortest hypoxic event occurred in 2014,
lasting 2 days. The longest hypoxic event was 71 days in 1989. Over the full period of
record (1987-2014), the average annual maximum duration was 31 days. Other estuaries
identified in the 2008 ISA where hypoxic conditions occur included Chesapeake Bay and
the Pamlico Estuary in North Carolina (U.S. EPA. 2008a). Murphy et al. (2011) analyzed
a 60-year record of Chesapeake Bay hypoxic zone volumes and found a slight but
significant decreasing trend in late summer hypoxia, which aligns with the decrease in N
loading due to management controls. This analysis also showed a correlation between N
loading and the duration of summer hypoxia events (Murphy et al.. 2011). In a modeling
study in Chesapeake Bay using analysis of monitoring data, nutrient loading was
identified as the main mechanism driving interannual hypoxia variabi 1 it\ (Li et al..
2016b). In upwelling regions such as along the West Coast, hypoxic events driven by
10-31
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
upwelling can be advected into estuaries from global ocean circulation rather than
anthropogic nutrient additions (Brown and Power. 2011; Brown and Ozretich. 2009).
10.2.5. Submerged Aquatic Vegetation
SAV, rooted vascular plants that grow to the surface but do not emerge from the water, is
important to the quality of coastal ecosystems because it provides habitat for a variety of
aquatic organisms, serves as nursery grounds for estuarine invertebrates and fish, absorbs
excess nutrients, and traps sediments (U.S. EPA. 2008a; Handlev et al.. 2007; Lefcheck
ct al.). Recently, the presence of seagrass beds was linked to decreased bacterial
pathogens of humans, fishes, and invertebrates in the water column and a lower incidence
of disease in adjacent coral reefs (Lamb et al.. 2017). The loss of SAV can, therefore,
have a cascade of effects on other ecosystem characteristics and ecosystem services.
Water clarity is important for photosynthesis of SAV (Handlev et al.. 2007). Water
quality changes associated with excess N nutrient inputs, such as excessive algal growth
and hypoxia can impact SAV extent. For example, declines of Johnson's sea grass
(Halophila johnsonii) which is currently listed as threatened under the U.S. Endangered
Species Act has been linked to eutrophication of coastal waters, specifically low DO and
algal blooms that alter habitat by covering up substate (Hernandez et al.. 2016). Elevated
levels of N tend to increase epiphytes on the surface of SAV, in the absence of other
limiting factors, contributing to declines in seagrass biomass, shoot density, percentage
cover, production, and growth (G.Nelson. 2017; Nelson. 2017).
Seagrass loss is occurring globally with nutrient enrichment as a major driving factor
contributing to declines in SAV coverage (Latimer and Rcgo. 2010; Wavcott et al..
2009). SAV is included as a biological indicator for estuarine condition for U.S. coastal
waters in ASSETS-ECI in the NEEA and in U.S. EPA's Report on the Environment (U.S.
EPA. 2016k; Bricker et al.. 2007). At the time of the 2008 ISA, estimates of historical
losses of SAV and declines in habitat were available for some coastal regions of the U.S.;
although, there were few data documenting the long-term response of SAV to N loading.
In Waquoit Bay, MA, (Valiela et al.. 1992) reported a strong negative relationship
between modeled N loading and measured eelgrass (Zostera marina) area based on
measurements of eelgrass coverage from 1951 to 1990. In the NEEA report, only a small
fraction of the estuary systems evaluated reported high severity of SAV loss from the
1990s to 2004 (Bricker et al.. 2007). in part, because historical losses wiped out
seagrasses in many estuaries, and recent changes do not seem of high severity.
Since the 2008 ISA, additional studies are available on the relationship between N
loading and SAV abundance, including the development of thresholds of response to N
June 2018
10-32
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
loading in seagrasses (Table 10-4). Seagrass dieback in Snug Harbor Cape Cod, MA in
2010 was linked to N loading, which stimulated growth of epiphytes on the seagrasses
(Howarth et al.. 2014). Orth et al. (2010) observed a consistent negative correlation
between SAV abundance and N loading based on water quality data in Chesapeake Bay
from 1984-2006. In the Potomac River, a major tributary to Chesapeake Bay, a reduction
in total N from point and nonpoint sources was significantly correlated to increased SAV
abundance and diversity using field data from 1990 to 2007 (Ruhl and Rvbicki. 2010).
Similarly, the reduction of N and P input from point sources into Mattawoman Creek, a
tributary to Chesapeake Bay, led to large increases in SAV coverage and density
(Bovnton et al.. 2014). Benson et al. (2013) identified a tidal-averaged total N
concentration of <0.34 mg/L as a threshold for healthy eelgrass in a survey of
19 Massachusetts estuaries. Eelgrass coverage decreases markedly in shallow estuaries in
New England with N loading rates >100 kg N/ha/yr (based on nitrogen loading model
[NLM] estimates from wastewater, fertilizer, and atmospheric deposition inputs).
Loading rates above 50 kg/ha/yr are likely to impact habitat extent (Latimer and Rcgo.
2010). These ranges were found to be comparable to loading thresholds identified in
other East Coast estuaries (Table 10-4). In a modeling study using NLM applied to
small- to medium-sized estuaries of southern New England, direct atmospheric
deposition to the water surface made up an average of 37% of the N input, while indirect
deposition via the watershed averaged 16% of N loading. The percentage, however,
varied widely for individual estuaries (Latimer and Charpentier. 2010).
10-33
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
Table 10-4 Nitrogen loading thresholds from multiple watershed sources versus
eelgrass loss.
Loading Threshold (kg/ha/yr)
Description
>20
100
•	Areal cover of eelgrass sharply reduced
•	Meadows disappeared rCape Cod estuaries, n = 10: (Bowen et
al.. 2007)1
-30
>60
•	Substantial eelgrass loss (80-96% of bed area)
•	Total disappearance TCape Cod estuaries, n = 7: (Hauxwell et al.,
2003)]
>64a
• Threshold based on nonparametric change-point analysis [95%
probability of change; Chesapeake Bay estuaries, n = 101; (Li et
al., 2007)1
>52
• Threshold based on nonparametric change-point analysis [95%
probability of chanae: New Enaland estuaries, n = 57: (Latimer
and Reqo. 2010)1
Consensus of Literature
Percentage of Eelgrass Area Loss (n = 57)
Mean Median 25th Percentile 75th Percentile
% %
<50
62 73 39 78
51-99
00
CO
00
CD
00
K)
CD
00
>100
93 100 95 100
aThis only includes point source inputs.
Source: Latimer and Reqo (2010).
The extent of SAV is stable or increasing in some coastal areas of the U.S. In Chesapeake
Bay, SAV coverage has increased from 41,000 acres (16,600 hectares) in 1978 to a peak
of 97,000 acres (39,250 hectares) in 2016 based on data collected by the Virginia Institute
of Marine Science KVIMS. 2016): Figure 10-61 as reported in U.S. EPA's Report on the
Environment (U.S. EPA. 2016k). SAV acreage has fluctuated in the bay since 2002,
covering an estimated 60,000 acres in 2013 then increasing in the most recent surveys
conducted in 2015 and 2016. This estimate is based on black-and-white, l:24,000-scale
aerial photographs of the Chesapeake Bay. Aerial monitoring first occurred in 1978 and
has occurred annually since 1984 (except during 1988). The SAV survey targets SAV
species, which as a group are sensitive to disturbance, particularly from eutrophication
and associated reductions in light availability. Pilots follow fixed flight routes to
comprehensively photograph all tidal shallow water areas of the bay and its tidal
June 2018
10-34
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
tributaries. Orth et al. (2010) reported a strong correlation between tributaries where
nutrient reductions have occurred and increases in seagrass abundance; however, further
reductions are necessary to meet SAV restoration targets. (Lcfchcck et al.) project a
catastrophic 95% loss of eelgrass Z. marina in Chesapeake Bay (and the associated
ecosystem services provided by this habitat type) in the next 30 years given the
conservative expectation of 2°C increase in temperature and continued trajectory of 40%
decline in water clarity due to the additive effects of these stressors especially on shallow
eelgrass beds. They report a 29% decline in eelgrass area in Chesapeake Bay since 1991
using high-resolution aerial imagery and water quality data. The authors note that
eelgrass abundance has increased in recent years but that recovery is limited to shallow,
nearshore areas and represents only a fraction of pre-1970s distribution of this species.
Despite extensive restoration efforts in the Chesapeake Bay watershed, water quality
indicators (chlorophyll a, DO, and Secchi depth) and biotic metrics of ecological
condition have shown little improvement during a 23-year period (Williams et al.. 2010)
The one exception to the pattern of no improvement in water quality was an observed
increase in the amount of SAV. Overall, water quality has decreased and chlorophyll a
levels have increased since 1986.
In Tampa Bay, FL, (see Appendix 16.4) data on seagrass (primarily Thalassia
testudinum) extent is available as far back as the 1950s. In the 1970s and early 1980s, the
bay exhibited signs of increasing eutrophication including loss of seagrasses. Seagrass
coverage has recovered to relatively predisturbed conditions (approx. 15,380 ha)
following N controls implemented in the mid 1980s (Greening et al.. 2014; Greening et
al.. 2011). Although the nutrient-load hypothesis of (Brauer et al.. 2012) suggests that
algae can outcompete or block out light for other macrophytes under high or variable
nutrient loads, this is not always the case. For example, in Yaquina Bay in Oregon,
distribution of eelgrass has remained stable despite high nutrient loads and algal blooms
(Kaldv. 2014). and seagrass populations in Puget Sound, WA have also remained
relatively stable for the last 40 years (Shelton et al.. 2017).
In other U.S. coastal areas, SAV coverage has declined. SAV coverage in the coastal
lagoon of Maryland and Virginia showed an increase with time until the early 2000s, and
then a decrease in the past decade, likely due to increasing anthropogenic nutrient inputs
that were accelerated by increased freshwater flow over the same time period (Glibert et
al.. 2014). In an analysis of 12 years of data for eelgrass areal coverage along the
Massachusetts coastline, there was an overall decrease in seagrass abundance although
the amount of change was highly variable between individual embayments (Costello and
Kenworthv. 2011).
10-35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Exhibit 1. Extent of submerged aquatic vegetation (SAV) in the Chesapeake
Bay, 1978-2016
Tj
1 50
100
5
Cn 50
Estimated additional acreage
' Mapped acreage
1978 1982 19S6 1990 1994 1 99B 2002 200S 2010 2014
Year
There were partial Bay-wide surveys from 1 979 to 1 983, and no survey in 1 9S8.
Information on the statistical significance of the trend in this exhibit is not currently available. For
more information about uncertainty, variability, and statistical analysis, view the technical
documentation for this indicator.
Data source: Chesapeake Bay Program, 201 7
SAV = submerged aquatic vegetation.
Source: U.S. EPA (2016k).
Figure 10-6 Extent of submerged aquatic vegetation in the Chesapeake Bay
1978-2016.
SAV is often at a competitive disadvantage under N enriched conditions due to the fast
growth of opportunistic macroalgae that preferentially take up NH4+ and can block light
from seagrass beds (Abrcu et al.. 2011). Eelgrass from the Pacific Northwest exhibited
increased growth rates with increasing NH4+ concentrations, but growth rate was not
related to NO;, concentration (Kaldv. 2014). Temperature was found to significantly
affect the response of eelgrass to nutrients, but the effects were not synergistic. A
significant increase in leaf length and decreases in shoot density and aboveground and
belowground biomass in eelgrass was observed concurrent with increased shading by
algae in field studies of 12 estuaries in Atlantic Canada (Schmidt et al.. 2012). In the
same study, C and N storage (two ecosystem services provided by eelgrass) declined with
increasing eutrophication.
A modeling framework to link variable freshwater inputs with seagrass (Hctlodule
wrightii, Thalassia testudinum) biomass and timing of growth and chlorophyll a was
June 2018
10-36
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
applied to the Caloosahatchee River estuary in southwest Florida (Buzzclli ct al.. 2014).
The model predicted organic N in the upper estuary, and chlorophyll a in time and space
reasonably well; however, there was more variation in NH44" while NOx (nitrate-nitrite)
was proportional to freshwater inflow. Overall, the model reflected variations in seagrass
biomass, although timing and growth did not match the variability of field observations.
A spatial model of coastal Australia was used to relate N loading from land to the extent
of SAV (Fcrnandcs et al.. 2015). The largest N plume was associated with discharges
from an industrialized estuary and a wastewater treatment plants. The location and size of
the N plumes changed with seasonal influences. The results of spatial model analysis
comparing the plume to seagrass distribution obtained from video surveillance showed
that dense seagrass meadows only occurred in areas that were unaffected by N plumes,
regardless of the seasonal influences on the plumes.
10.2.6. Indices of Estuarine Condition
Biological and chemical indicators have been used by the states to develop numeric
nutrient criteria for estuaries (Appendix 7.2.10). Indicators may also be combined into an
overall condition rating to measure ecosystem function, structure and processes in a
standardized approach (U.S. EPA. 2016g; Boria et al.. 2012; U.S. EPA. 2012b; Devlin et
al.. 2011; Bricker et al.. 2007). Several assessment frameworks for eutrophic condition
have been developed in the U.S. (e.g., U.S. EPA's NCCA and the NOAA NEEA
ASSETS-ECI) and other countries (e.g., Trophic State Index [TRIX], Institutfrangais de
recherche pour Vexploitation de la mer [IFREMER], transitional water quality index
[TWQI]); however, the applicability of a specific framework to areas outside of the
region where they were originally developed may be limited (McLaughlin et al.. 2014;
Boria etal.. 2012). Within the same estuary, results of assessment of eutrophic condition
may vary depending on the framework used for evaluation and the associated chemical
and biological indicators (McLaughlin et al.. 2014; Garmendia et al.. 2012; Devlin et al..
2011). In a comparison of methods to assess nutrient enrichment impacts, Devlin et al.
(2011) suggested that indices incorporating annual data, frequency of occurrence, spatial
coverage, secondary biological indicators, and a multicategory rating scale are more
robust and representative. Two of these methods that have been applied to U.S. waters
are ASSETS-NCI (Bricker et al.. 2007) and NCCA (U.S. EPA. 2016g. 2012b).
ASSETS-NCI is only applied to nutrients and focuses on response indicators rather than
chemical indicators, while the NCCA can be used for assessing other stressors to coastal
areas and integrates both chemical and biological data.
The ASSETS ECI in the NEEA (Bricker et al.. 2007) was used to estimate the likelihood
that an estuary is experiencing or will experience eutrophication based on five ecological
June 2018
10-37
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
indicators: chlorophyll a, macroalgae, DO, nuisance/toxic algal blooms, and SAV
[Figure 10-3; (Bricker et al.. 2007)1. ASSETS uses the frequency and spatial extent of
algal blooms combined with the 90th percentile of annual values for chlorophyll a
(Table 10-2). Estuaries are divided into salinity zones, and ratings are combined as an
area-weighted sum. Results from the NEEA were included in the 2008 ISA, and the
biological indicators are the same in both reports for nutrient effects in estuarine systems.
The NCCA reports represent collaboration between U.S. EPA, NOAA, U.S. Fish and
Wildlife Service, and coastal state agencies. The most recent sampling period was 2010
(U.S. EPA. 2016g). NCCAs use chlorophyll a DO, and three additional indicators (DIN,
DIP, water clarity) to determine a water quality index (U.S. EPA. 2016g. 2012b). The
NCCA include data on water quality, sediment quality, benthic community composition,
and fish tissue contaminants to determine the overall condition of the nation's coastal
waters. In the most recent NCCA report, water quality was rated good in 36% of coastal
and Great Lakes nearshore waters, fair in 48%, and poor in 14%, based on measures of
the eutrophication parameters that make up the water quality index (P, N, water clarity,
chlorophyll a, and DO concentrations).
Additional indices applied to U.S. waters include biological indicators of estuarine
condition. (Fertig et al.. 2014) describe a Eutrophication Index applied to Barnegat
Bay-Little Egg Harbor Estuary, NJ using weighted indicators of water quality
(temperature, DO, TN, TP), light availability (chlorophyll a, total suspended solids,
Secchi depth, macroalgae percentage cover, percentage surface light, epiphyte biomass),
and seagrass (Zostera spp.) response (aboveground biomass, belowground biomass,
density, percentage cover and length). The biological condition gradient conceptual
framework (Davies and Jackson. 2006) was applied to Greenwich Bay, RI to assess
estuarine habitat overtime (Shumchenia etal.. 2015). Biological indicators included
seagrass extent, benthic community, primary productivity, and shellfish. In a regional
survey of 23 estuaries in southern California, 78% of estuaries were rated "moderate" or
"worse" based on macroalgae, 39% based on phytoplankton, and 63% based on DO using
the European Union Water Framework Directive (McLaughlin et al.. 2014). With the
ASSETS framework, 53% of surveyed areas were identified as impaired. The variability
in categorizing estuarine condition was influenced by spatial and temporal scales as well
as which indicators and thresholds were selected. Additional indices used in other
countries for describing estuarine condition are reported in Zaldivar et al. (2008). Boriaet
al. (2008). Boria et al. (2012). Devlin etal. (2011). Garmendia et al. (2012). and
Andersen et al. (2014).
A modeling indicator for ecosystem response to N uptake, which includes a measure of
the loss of species richness, was used to create a marine eutrophication Ecosystem
10-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Damage indicator (meED) (Cosmc et al.. 2017). The indicator is based on the loss of
species richness caused by hypoxia, which is in turn caused by eutrophication. The meED
ecosystem damage indicator itself indicates the potential impact of eutrophication in the
receiving habitat based on species density estimates. Cosme et al. (2017) support the
meED for inclusion in Life Cycle Impact Assessments (LCIA) when characterizing N
emissions. Their paper describes the calculation of the meED indicator for 66 large
marine ecosystems and maps risk of ecosystem damage due to eutrophication based on
this indicator.
10.3. Effects on Biodiversity
Increased N loading to coastal areas can lead to shifts in community composition,
reduced biodiversity, and mortality of biota. Biodiversity is important for ecosystem
stability and function, including provision of ecosystem services (Chapter 1.2.2.4).
Evidence for impacts to biodiversity include paleontological evidence (Appendix 10.3.1).
altered phytoplankton community composition (Appendix 10.3.2). responses of
phytoplankton to reduced versus oxidized N (Appendix 10.3.3). bacteria/archaea
diversity (Appendix 10.3.4). benthic diversity (Appendix 10.3.5). and fish diversity
(Appendix 10.3.6). The form of N supplied can significantly affect phytoplankton
community composition in estuarine and marine environments (Glibert et al.. 2016; Paerl
et al.. 2000; Stolte et al.. 1994). In hypoxic areas, mortality of benthic biota and
avoidance of I0W-O2 conditions by mobile organisms lead to changes in biodiversity and
loss of biomass (Diaz and Rosenberg. 2008). Energy transfer through the food web can
also be altered by a decrease in predators and an increase in microbes from oxygenated
areas to anoxic zones.
10.3.1. Paleontological Diversity
Sediment records from Chesapeake Bay showed alterations in producers and consumers
correlated to land use change in the watershed (Sowers and Brush. 2014; Brush. 2009). In
this estuary, diatom community structure has shown a steady decrease in overall diversity
since 1760 as estimated by sediment core analysis (Cooper and Brush. 1993). Diatom
populations began to shift from benthic to planktonic in the early 1900s corresponding to
the increased N flux and higher sediment inputs [which decreased light penetration;
(Brush. 2009)1. Additional sediment cores from two distinct locations in the estuary show
a shift to an estuarine food web that is predominately planktonic (Sowers and Brush.
2014). An increase in the abundance of the formaminifera Ammobaculites spp. and a
decrease in the abundance of the polychaete Nereis spp. were observed along with the
June 2018
10-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
change in primary producers. Paleontological evidence from England shows significant
changes in benthic species composition and ecosystem functioning between periods of
extremely low O2 (Caswell and Frid. 2013). The evidence suggests that normal benthic
functions are maintained during early hypoxia, but these functions collapse once
thresholds of severely low O2 are reached.
10.3.2. Phytoplankton Diversity
As reported in the 2008 ISA, excess N can contribute to changes in phytoplankton
species composition (U.S. EPA. 2008a'). High loadings of N and P can also increase the
potential for Si limitation, with associated changes in diatoms. Such changes to the
phytoplankton community and functional groups (e.g., diatoms, dinoflagellates,
cryptomonads, cyanobacteria, chlorophytes) can also affect higher trophic levels because
these organisms are the base of the estuarine food web. For example, if the nutrient mix
favors species that are not readily grazed (e.g., cyanobacteria, dinoflagellates), trophic
transfer will be poor, and relatively large amounts of unconsumed algal biomass will
settle to the bottom, which could stimulate decomposition, O2 consumption, and the
potential for hypoxia (Paerl et al.. 2003). Changes in phytoplankton species abundance
and diversity have been further documented through in situ bioassay experiments, such as
the results reported by (Paerl et al.. 2003) for the Neuse River estuary, North Carolina.
Effects were species specific and varied dramatically depending on whether, and in what
form, N was added. The findings illustrate the potential impacts of N additions on
phytoplankton.
Several studies published since the 2008 ISA use shifts in diatoms as a measure of N
enrichment effects on species diversity. In microcosm studies from Raritan Bay, centric
and chain-forming diatoms resulted from enriched NO3 concentrations, differing from
the pennate diatoms and green flagellates that accompanied ambient NO3 concentrations
(Rothenberger and Calomeni. 2016). The Si:N ratio was identified as an important factor
governing the phytoplankton community dynamics in the bay. Phytoplankton community
structure was altered by eutrophication in the Skidaway River estuary, Georgia, as
evidenced by a shift away from diatoms and towards nonsilicate nanoplankton (Verity
and Borkman. 2010). There was a strong relationship between diatom taxon distribution
and TN concentrations in Charlotte Harbor, FL (Nodine and Gaiser. 2014). Differences
were noted in the phytoplankton community structure during the time periods before,
during, and after eutrophication in the Black Sea (Mikaelvan et al.. 2013). Diatoms and
dinoflagellates especially were found to be dependent on high N and N:P ratios. While
nutrient ratios are known to influence phytoplankton community structure, Davidson et
June 2018
10-40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
al. (2012) pointed out that these nutrient ratios are important only when at least one
nutrient is low enough to limit growth.
Shifts in phytoplankton community structure are known to occur in estuaries due to N
enrichment, but climatic changes may at times outweigh the impacts of eutrophication
I (Pacrl et al.. 2010); Appendix 10.1.4.11. Thus, physical changes caused by climate
change, such as water temperature, stratification, circulation, and hydrologic variability,
must be taken into account when modeling phytoplankton community responses to N
enrichment (Pacrl et al.. 2014). Likewise, Glibert et al. (2014) noted changes in
phytoplankton community composition along with increased anthropogenic N input and a
switch to long-term wet conditions in the coastal lagoons area of Maryland and Virginia.
A 3-year data set from Raritan Bay indicates that both climatic conditions and nutrient
concentrations affected phytoplankton composition from 2010-2012 (Rothcnbcrgcr et al..
2014). The abundance of flagellates in the bay over time reflects a shift in dominance
away from diatoms that is characteristic of eutrophic systems. Location and surrounding
land cover, as well as the type of N input, are important factors in the response of the
phytoplankton community to N addition (Reed et al.. 2016). Phytoplankton community
composition in four tidally influenced sites along a gradient from highly developed to
undeveloped (natural) in coastal SC were influenced by surrounding land cover and the
form of N, with cyanobacteria, dinoflagellates, and other potentially HAB-forming
species most often found at the more developed sites.
10.3.3. Diversity of Phytoplankton Under Different Forms of Nitrogen (Reduced
vs. Oxidized)
Reduced forms of atmospheric N are increasing relative to oxidized N in the U.S.
(Appendix 10.1.2) and play an increasingly important role in eutrophication and HAB
dynamics in coastal areas. Specific phytoplankton functional groups have a preference for
NH4+ over NOs (Appendix 10.2.2). perhaps leading to selective stimulation of primary
production, especially in light-limited estuarine and coastal waters (Glibert et al.. 2016;
Paerl et al.. 2000; Stolte et al.. 1994). Increasing loads of NH3+/NH4+ have been linked to
the expansion of HABs (Glibert et al.. 2016; Esparza et al.. 2014; Altman and Paerl.
2012; Blomqvist et al.. 1994). For example, in San Francisco Bay, a comparison of N
isotopes in cells of the cyanobacterial HAB species Microcystis aeruginosa with N in
rivers flowing into the bay indicated the primary source of N that supported the bloom
formation was likely NH4+ (Lehman et al.. 2015). Phytoplankton response to the form of
N supplied (Appendix 10.2.2. and Table 10-3) could lead to altered phytoplankton
growth and community composition and have cascading effects on trophic structure and
biogeochemical cycling (Glibert et al.. 2016; Paerl et al.. 2000). Cyanobacteria, and many
June 2018
10-41
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
chlorophytes and dinoflagellates may be better adapted to the use of NH/, while diatoms
generally thrive in environments with oxidized forms of N such as NO3 . Figure 10-7
summarizes the dominant functional phytoplankton groups in primarily reduced N and in
primarily oxidized N conditions.
Field studies indicate that HAB-forming dinoflagellates are often associated with
enrichment of reduced N. In Maryland and Virginia coastal bays, Glibert et al. (2014)
observed a regional shift in phytoplankton community composition to species that do
well in reduced N conditions. Phytoplankton community dynamics varied with the form
of N in nutrient enrichment experiments using water from the New River estuary, NC
(Altman and Paerl. 2012). Photopigment analysis used to identify and quantify taxonomic
groups revealed that the addition of riverine dissolved organic nitrogen (DON) promoted
dinoflagellates, chloropytes, and myxoxanthophyll (cyanobacteria) while zeaxanthin
(cyanobacteria) was most frequently detected with inorganic N. In the Neuse River
estuary, NC where NH44" concentrations have increased over time, the abundance of
aphidophytes (including the potentially toxic dinoflagellate species H. akashiwo),
haptophytes, chlorophytes, and the bloom-forming dinoflagellate Heterocapsa rotundata
have also increased (Rothenberger et al.. 2009). Incubation experiments carried out in the
Neuse River estuary in spring, early summer, and late summer to test how the growth and
community composition of phytoplankton responded to N availability and specific form
of N (Cira et al.. 2016). Phytoplankton community composition varied by season, and
different taxa were found to be able to uptake both urea and NO3 when conditions were
limiting, although the results varied by taxa. Seasonal experiments suggested that N was
not the only factor controlling phytoplankton growth in the spring, as the experiment
conducted in March showed that only growth rates of fucoxanthin-containing (brown
algae) species were stimulated by N addition. In field studies along a gradient from
highly developed to undeveloped (natural) study sites in the southeastern U.S.,
phytoplankton communities at the more developed sites showed higher biomass and
growth rates with N (particularly urea) additions (Reed et al.. 2016). In addition,
cyanobacteria, dinoflagellates, and other potentially HAB-forming species were more
often found at the more developed sites.
10-42
DRAFT: Do Not Cite or Quote

-------
diatoms
crypto-
phytes
dirio
flagellates
pico-
tyanobacteria
S	I
5=|

C. u	•Ł*
M_	J2
—j O	'."1J
5 urn


Source: Gilbert et al. (2016).
Figure 10-7 Summary conceptual schematic illustrating the effect of changes
in the proportion of Nm+ and NO3" in the loads of N provided to a
natural system. When NtVis the dominant form, and when waters
are warmer, flagellates, cyanobacteria, and chlorophytes among
other classes may proliferate, leading to overall productivity
dominated by the small size class of algae (e.g., <5 |jm). In
contrast, when NOs" is the dominant form provided, especially
under cooler water conditions, diatoms more likely dominate, and
their overall production will be more likely dominated by cells of a
larger size class (e.g., >5 pm). Moreover, chlorophyll a yield and
total production may be higher than under the NH4+enrichment
condition.
June 2018
10-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
In mesocosm studies in the P-limited Florida Bay ecosystem, chl a concentrations
(phytoplankton biomass) increased with N + P treatment but did not increase with only N
enrichment. However, N only enrichment (especially in the form of NH44") did change the
phytoplankton assemblage in the direction of more picocyanobacteria (Shangguan ct al..
2017). A shift of phytoplankton biomass toward diatoms making up a larger portion of
the total biomass was observed when N in the form of NO3 was added, either alone or in
combination with P. Using chlorophytes and diatoms from Suisun Bay, CA grown under
controlled laboratory conditions Berg et al. (2017) found that the four diatoms species
tested grew faster on NH44" than NOs suggesting that diatoms were not under a
competitive disadvantage under high NH44". Only the chlorophyte R. planktonicus grew
significantly faster on NO3 . None of the diatoms tested in controlled lab conditions were
sensitive to NH44" at concentrations that would be found in the local environment. The
results show that the responses are highly species specific in tolerance of NH4+.
Not all studies have found variation in algal response with the form of N. Richardson et
al. (2001) examined the effects of different forms of N application (NO3 . NH4+, urea) on
the structure and function of estuarine phytoplankton communities in mesocosm
experiments in the Neuse River estuary, NC. Even though NH4+ is more readily taken up
by phytoplankton in this estuary than is NO3 (Twomev et al.. 2005). the results of the
Richardson et al. (2001) study suggested that phytoplankton community structure was
determined more by the hydrodynamics of the system than by the form of N available for
growth. Twomev et al. (2005) measured Neuse River estuary phytoplankton uptake rates
of NH4+, NO3 . and urea. NH44" was the dominant form of N taken up, contributing about
half of the total N uptake throughout the estuary. Uptake varied spatially; in particular,
N03 uptake declined from 33% of the total uptake in the upper estuary to 11 and 16% in
the middle and lower estuary, respectively. Urea uptake contributed 45 and 37% of the
total N uptake in the middle and lower estuary, showing the importance of regenerated N
for fueling phytoplankton productivity in the mesohaline sections of the estuary.
Therefore, N budgets based only on inorganic forms may seriously underestimate the
total phytoplankton uptake (Twomev et al.. 2005).
10.3.4. Diversity of Bacteria, Archaea, and Microzooplankton
Since the 2008 ISA, a few studies have explored ammonia-oxidizing bacteria (AOB) and
ammonia-oxidizing archaea (AOA) community responses in relation to N in coastal
waters. These organisms play a key role in N cycling and the AOA were described
relatively recently (Konneke et al.. 2005). Bouskill et al. (2012) examined the distribution
and abundance of AOA and AOB across a variety of aquatic environments, including two
sites at different salinity levels in Chesapeake Bay. The AOB were dominate in the
June 2018
10-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
freshwater end of the continuum, whereas AOA were more common in open ocean areas.
The abundance of both groups were correlated with the concentration of NH4+ in the
water, especially relative abundance of AOB. A similar pattern was observed in the
Sacramento-San Joaquin Delta where AOB were more abundant in the NH4+-rich
Sacramento River compared to the San Joaquin River and Suisun Bay where AOA were
prevalent (Damashck et al.. 2015). Community structure of benthic ammonia oxidizers
differed across the region and appeared to be related to nutrient inputs. In the Puget
Sound estuary, WA, AOA were generally more abundant, however, AOB increased
relative to AOA during periods of high NH4+ (Urakawa et al.. 2014). Community
structure of plantonic ciliates was found to be significantly related to the spatial
distribution of NO;, -N concentrations across a gradient of sites in Jiaozhou Bay, China,
suggesting the potential applicability of plantonic ciliate functional groups as a water
quality indicator (Xu et al.. 2016a).
10.3.5. Benthic Diversity
Invertebrates associated with bottom substrates of coastal systems are useful indicators of
biological change for impacted waters. Several indices based on benthic invertebrates,
macroinvertebrates, or macrophytes have been used to assess biological condition in
response to the European Water Framework Directive (Andersen et al.. 2014; Zaldivar et
al.. 2008). A few of these indices have been tested in U.S. waters. Boriaetal. (2008)
compared the Benthic Index of Biotic Integrity (B-IBI), the AZTI Marine Biotic Index
(AMBI) and its multivariate extension the M-AMBI in Chesapeake Bay. There was
relative agreement between methodologies, and differences were related to spatial
variability and habitat type. Comparable regional benthic indices for the
Northeast/Acadian, northeast Virginian, southeast Virginian, southeast Carolinian, and
Gulf Louisianan coasts were developed for the U.S. EPA's coastal assessments (NCCA)
(U.S. EPA. 2016g).
N enrichment has been shown to have significant but complex effects on
suspension-feeding macroinvertebrates in estuaries. Shellfish species can respond
positively to increased N loading and high phytoplankton biomass levels due to increased
quantity and quality of food particles (Carmichael et al.. 2004). A recent study in Long
Island's Peconic Estuary suggested that eutrophication affects shellfish through changes
in the quality of food and not the quantity (Wall et al.. 2013). Select species (oysters [C.
virginica] and clams [Mercenaria mercenaria]) in eutrophic areas experienced enhanced
growth rates that were strongly correlated with high densities of autotrophic
nanoflagellates and centric diatoms. Other species (scallops [Argopecten irradians| and
slipper limpets) suffered negative effects and grew at the slowest rate at the most
June 2018
10-45
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
eutrophic sites. Bay scallop growth was negatively correlated with densities of
dinoflagellates, which were more abundant at the most eutrophic site (p < 0.001). In
rocky shore environments in Ireland, similar to those of the U.S. northeastern Atlantic
coast, N enriched sites near industrial and sewage outfalls have been shown to have
reduced total abundance and number of molluscan taxa and altered community
composition compared with control sites (Atalah and Crowe. 2012). Nitrite and NH3
levels explained the highest percentages of variation in the molluscan assemblage
structure (13 and 12.5%, respectively), owing in large part to the absence of three rare
species at contaminated sites. The observed shift in community structure caused by
different levels of molluscan tolerance to N enriched conditions has been suggested for
use as a biological indicator of eutrophication (Atalah and Crowe. 2012). The role of
benthic invertebrates in coastal N cycling and the use of shellfish for coastal N
remediation is discussed in Appendix 7.2.6.11.
10.3.6. Fish Diversity
A few studies have recently reported effects on fish biodiversity in nutrient-impacted
estuaries. Comparison of trophic organization of fish in estuarine reaches of nine rivers in
Victoria, Australia showed that inorganic N loading was an important factor explaining
trophic diversity of fish assemblages. There was greater trophic diversity of fish
assemblages with mid to high inorganic N loading suggesting that DIN influences not
only estuary primary productivity but is transmitted upward through the food chain
(Warn etal.. 2016). In Prince Edward Island estuaries, adult and young-of-the-year
abundance of pollution-tolerant mummichogs (Fundulus heteroclitus) increased in
estuaries with greater intensity of N loading from agricultural land use (Finlcv ct al..
2013). In this study, eutrophication was associated with changes in habitat variables,
which were linked to mummichog abundance. In an estuary in the southern Gulf of St.
Lawrence in Canada, loss of eelgrass was linked to declines in fish diversity, but the loss
did not change the positions of organisms within the food chain (Schein et al.. 2013;
Schein et al.. 2012).
10.3.7. Trophic Interactions
Altered trophic transfer following nutrient-associated changes in phytoplankton
community composition were reported in the 2008 ISA. Phytoplankton that are not
readily grazed, such as cyanobacteria and dinoflagellates, are not transferred to higher
trophic levels as efficiently as diatoms [more readily grazed by zooplankton which are
then consumed by fish, (Pacrl et al.. 2003)1. Since the 2008 ISA, studies have further
June 2018
10-46
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
characterized the effects of nutrient enrichment on trophic interactions. In the Skidaway
River estuary, GA, increasing nutrient concentrations and changes in nutrient ratios led to
increases in eukaryotic organisms (heterotrophic and mixotrophic flagellates,
dinoflagellates, ciliates, and metazoan zooplankton) but not diatoms, potentially altering
food web predator-prey relationships (Verity and Borkman. 2010). In an experimental
study in Chesapeake Bay, Reynolds et al. (2014) demonstrated that a reduction of
crustacean grazers controlling algal growth on seagrasses resulted in 65% reduced
seagrass biomass and nearly sixfold increased epiphytic algae. In experimental plots
where grazers were removed, aboveground eelgrass biomass was reduced following
fertilization. In the presence of grazers, eelgrass biomass increased with fertilization.
When predators were excluded, mesograzers were able to limit ephiphyte growth even
when nutrients were added. Based on these findings, the researchers suggested that it is
unlikely that reducing nutrient pollution alone would restore seagrass meadows where
alterations to food webs have already reduced populations of algae-feeding mesograzers.
Similar results were reported in McSkimming et al. (2015) from Posidonia angustifolia
meadows in Australia where mesograzers responded to nutrient addition by increasing
grazing per capita, resulting in greater consumption of epiphytic algae.
A review of nutrient addition studies from the North Atlantic suggests that, overall, the
addition of nutrients caused the biomass of opportunistic macroalgae to approximately
double. Higher numbers of midlevel predators has the same effect because they lead to a
smaller population of grazers eating the macroalgae (Ostman et al.. 2016). Analysis of
review data indicated that the effect of midlevel predators on green macroalgae biomass
increased with eutrophication. Recovery of sea otter (Enhydra lutris) populations in
California appear to control algal growth on eelgrass by increased predation of crabs
(Cancer spp. and Pugettia producta) which, in turn, decrease predation of mesograzers
such as sea slugs (Phyllaplysia taylori) that feed on the algae (Hughes et al.. 2013).
Newer literature has provided evidence for complex interactions between eutrophication
and other stressors and subsequent effects on trophic interactions. For example, Burnell et
al. (2013) used mesocosms to demonstrate the effect of eutrophic conditions on
plant-herbivore grazing interactions between the sea urchin Amblypneustes pallidas and
the seagrass Amphibolis antarctica affected by warming and acidification. Nutrient
enrichment offset increased grazing associated with acidification and warming; however,
it did not fully counter the additive effects of these two stressors.
10.3.8. Models Linking Indicators to Nitrogen Enrichment
Process-oriented models such as nutrient-phytoplankton-zooplankton (NPZ) models are
used to predict the response of organisms such as HABs to various known and/or
June 2018
10-47
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
predicted nutrient conditions (Swancv et al.. 2008). A nutrient-driven phytoplankton
model developed by Scavia and Liu (2009) was expanded by Evans and Scavia (2013) to
test the sensitivity of the response of chlorophyll a levels and DO levels to N enrichment.
Results indicated that separate processes control chlorophyll a and DO sensitivity
(estuary flushing time and relative mixing depth, respectively), and that these sensitivities
vary among estuaries (Figure 10-8).
Using the European regional Seas Ecosystem Model for an area of the Japan Sea, Lee and
Yoo (2016) showed that the phytoplankton community there will shift to smaller
phytoplankton, and the food web structure will likely be altered if atmospheric deposition
continues to intensify in the region as predicted. For the study period 2001-2012, the
model estimated that the atmospheric N deposition in the Ulleung Basin would increase
the annual net primary production by an average of 4.58% (range from 3.77-10.58%).
10-48
DRAFT: Do Not Cite or Quote

-------
25
20
"—•—Blue Hill B
— O— Patuxent R
. —*— Penobscot B
—O— Potomac R
—X— St, Marys R and Cumberland S
6
u 10
Chi = chlorophyll; DO = dissolved oxygen; TN = total nitrogen.
The x-axis shows river total nitrogen load relative to the 1992 SPARROW total nitrogen load for each estuary. In the figure legend,
R = river, B = bay, and S = sound.
Source: Evans and Scavia (2013).
Figure 10-8 Forecasting curves for effects on total nitrogen loadings on
(A) chlorophyll and (B) dissolved oxygen (mean and 95%
confidence interval) for selected estuaries demonstrating the full
range of sensitivity to relative total nitrogen loading.
June 2018
10-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
10.4. Animal Behavior and Disease
In addition to changes in biological indicators (i.e., chlorophyll a, HABs, macroalgal
abundance, SAV) and altered species diversity there is increasing evidence for a role of N
in behavior and disease in coastal biota. These indirect responses may affect the fitness of
organisms inhabiting nutrient-enriched waters.
10.4.1. Behavior
Coastal biota exhibit behavioral responses to hypoxic conditions, including acclimation
or avoidancc(Townhi11 et al.. 2017; Levin et al.. 2009). Some fish have the ability for
aquatic surface respiration, a behavior in which the fish swims to the surface and is able
to take advantage of the relatively higher levels of DO in the water there. Dixon et al.
(2017) assessed four common species of shallow estuarine finfish for their reactions to
episodic hypoxic conditions and altered pH levels. Severe hypoxia induced surface
respiration behavior of three of the four fish species, but all were unaffected by diel
variations in pH.
Turbidity caused by excess algae in the water, a secondary effect of eutrophication, has
been shown in laboratory studies to potentially alter some fish behaviors, with
implications for mate selection and offspring survival rates. It is not known whether some
of these behavior changes will be adaptive or if they will have positive or negative effects
on the fish populations. For instance, when the strength of sexual selection on several
traits is relaxed, the relative importance of survival selection may increase (Candolin.
2009). The color contrast created by increased turbidity may lead to a heightened feeding
response in some fish larvae such as California yellowtail (Seriola lalandi), which
exhibited elevated growth and survival rates compared to larvae raised in clear
(oligotrophic) water (Stuart and Drawbridge. 2011).
Several studies of the three-spined sticklebacks (Gasterosteus aculeatus) indicate that
observed alterations of reproductive behaviors are linked to eutrophic conditions
(Figure 10-9). Laboratory manipulations suggest that turbidity caused by eutrophication
may affect mate selection and breeding success in this euryhaline fish species (Heuschele
et al.. 2012; Heuschele and Candolin. 2010). When given the choice between dense or
sparse algae, male sticklebacks preferred to nest in dense algae (which simulated
eutrophic conditions). The nests of those males were more likely to be parasitized by
other males; however, the males nesting in dense algae also acquired more eggs, and
thus, had a higher probability of reproduction (Heuschele and Candolin. 2010). This may
be because reduced visibility leads to reduced mate selectivity searching by females,
June 2018
10-50
DRAFT: Do Not Cite or Quote

-------
1
2
although it is unknown whether these behavior changes would have positive or negative
impacts on the population (Hcuschclc et al.. 2012).
Eutrophication
-	filamentous algae
-	water turbidity
Reproductive
success
I
Sexual selection
1
Population
demography
1
Survival selection
Source: Candoliri (2009).
Figure 10-9 The pathway of effects of eutrophication on different reproductive
behaviors and selection forces in Gasterosteus aculeatus.
3	Simulated turbidity has also been shown to impair visual mate choice in an eastern
4	Atlantic species of marine pipefish, Syngnathns typhis (Sundin et al.. 2010). A follow-up
5	study in the same species allowed other factors to occur, namely mating competition and
6	mate encounter rates which counteracted the effects of lack of visual cues and led to
7	increased sexual selection in turbid waters (Sundin et al.. 2017). In the same species, the
8	latency period between courting and copulation was prolonged in I0W-O2 conditions,
June 2018	10-51 DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
although other measures of reproductive behavior (total time spent courting, probability
of mating, dancing, male pouch-flap behavior) were unaffected (Sundin et al.. 2015).
Other types of fish mating behaviors are also affected by eutrophication. Male
sticklebacks engage in an opportunistic mating behavior called "sneaking" which is less
successful under more turbid (eutrophic) conditions (Vlicgcr and Candolin. 2009).
Eutrophication has so far been found to have conflicting impacts on nest building
behavior in male sticklebacks, with varying implications for the survival rate of
stickleback offspring. Male sticklebacks built smaller nests with wider openings during
eutrophic (algal bloom) conditions (Wong et al.. 2012). These males also took longer to
complete the nest than did males that built during nonbloom conditions. However, under
normal water conditions in a laboratory setting, male sticklebacks that had been collected
from eutrophic waters built their nests faster than those collected from noneutrophic
waters, although the nests did not differ in structure or composition between the two
groups (Tuomaincn and Candolin. 2013V In European sand gobies (Pomatoschistus
minutus), another species that exhibits paternal care of eggs, algal turbidity altered male
behaviors (Jarvenpaa and Lindstrom. 201IV Males spent more time away from the nest,
and exhibited less fanning of the eggs with their pectoral and tail fins. Goby egg survival
rates were actually found to be higher under turbid conditions. Although the mechanism
for this outcome is unknown, it was proposed that males may have increased other forms
of care in the presence of eutrophic conditions, leading to higher egg survival.
Under normal conditions, it is beneficial for sticklebacks to live in larger groups where
they gain protection from predators, are able to forage more efficiently, and are able to
find a mate more easily than in smaller groups. However, in eutrophic (turbid) water,
sticklebacks did not show a preference for joining a larger shoal as they did under normal
conditions (Fischer and Frommen. 2013). This lack of exhibited shoal preference under
turbid conditions could eventually reduce individual stickleback fitness. In addition,
sticklebacks changed shoals less frequently in turbid water than observed under normal
conditions, leading to a decrease in the transfer of important social information, such as
the location of foraging routes. Therefore, increased turbidity may reduce social learning
opportunities for sticklebacks and may, in turn, impair their foraging efficiency (Fischer
and Frommen. 2013).
10.4.2. Disease
Diel-cycling hypoxia in estuaries is a natural phenomenon that can be exacerbated by
increased biomass and productivity associated with anthropogenic nutrient loadings
(Tyler et al.. 2009). Tyler et al. (2009) observed that upper areas of four Delaware
June 2018
10-52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
estuaries experienced increased incidence of severe hypoxia (DO <2 mg/L) where
nutrient loading was greater. The duration and spatial extent of diel-cycling hypoxia in
the estuaries were primarily associated with water temperature, previous day's insolation,
hours of morning ebb tide, and daily stream flow. Oysters (C. virginica) exposed to
periods of diel-cycling hypoxia were demonstrated to have increased incidence and
progression of Perkinsus marinus infection (Dermo) in field and laboratory experiments
(Brcitburg et al.. 2015). Field experiments were conducted in Chesapeake Bay. The
authors suggest that the likely mechanism is negative effects of hypoxia on host immune
response.
Seagrass meadows were recently shown to decrease pathogens harmful to both human
health and marine organisms. In field studies adjacent to islands in Indonesia affected by
wastewater pollution, bacterial pathogens were reduced by up to 50% in seagrass
meadows and coral reefs adjacent to seagrass beds had reduced disease incidence (Lamb
et al.. 2017). Several recent studies have suggested that seagrasses may be more
susceptible to wasting disease caused by the marine slide mold (genus Labyrinthula)
under conditions of elevated NOs loading (Hughes et al.. 2017; Kaldv etal.. 2017).
Nutrient enrichment is one of several factors linked to increased disease susceptibility
and bleaching in corals (D'Angclo and Wiedenmann. 2014; Vega Thurber et al.. 2014;
Wiedenmann et al.. 2013). Coral bleaching occurs when the symbiotic relationship
between the coral host and microalgae (Zooxanthellae spp.) breaks down. Although most
research on nutrient loading on corals includes both N and P several studies have isolated
effects of N which impacts corals via distinct pathways from P (Shantz and Burkcpilc.
2014). Wiedenmann et al. (2013) reported that the coral-algae symbiosis can be
interrupted by elevated inorganic N concentration rather than both N and P. In a
metanalysis of nutrient impacts on corals, N reduced coral calcification 11% on average,
while increasing photosynthetic rate (Shantz and Burkepile. 2014). Increased DIN
decreases the temperature threshold at which coral bleaching occurs (Wooldridge. 2009).
The threatened status of staghorn coral (Acropora cervicornis) and elkhorn coral
(Acroporapalmata) under the U.S. Endangered Species Act has been linked to indirect N
pollution effects, specifically low DO and algal blooms that alter habitat, and other
non-nutrient stressors (Hernandez et al.. 2016).
10.5. Nutrient Enhanced Coastal Acidification
Since the 2008 ISA, several studies have suggested that the increased respiration caused
by N enrichment may exacerbate coastal ocean acidification through alteration of the
carbon cycle (Appendix 7.2.4). Dissolution of atmospheric anthropogenic carbon dioxide
June 2018
10-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
(CO2) into the ocean has caused increasing acidification of coastal waters, resulting in
long-term decreases in pH (Wallace et al.. 2014; Orr et al.. 2005). N enrichment is
expected to worsen this acidification because degradation of excess organic matter then
produces CO2 in the water column, which in turn can make the water more acidic
I Figure 10-10; (Wallace et al.. 2014; Sunda and Cai. 2012; Cai et al.. 2011c; Howarth et
al.. 2011)1. An important additional source of organic matter leading to overall declines
in pH is potentially allochthonous organic matter inputs which have been increasing in
many coastal watersheds from changing land use (Wilson et al.. 2016; Wetz and
Yoskowitz. 2013). Both N-driven eutrophication and anthropogenically enhanced
allochthonous organic matter loading operate simultaneously.
In additional to microbial degradation of organic matter, respiration of living algae and
seagrasses during the night can also drive acidification. In an N-enriched seagrass
ecosystem in Cape Cod, a very pronounced diel pattern in pH was observed with highest
acidity at dawn (Howarth et al.. 2014). The pH varied from below 7.0 at dawn to near
seawater levels of 8.1 by sunset. Acidification also can be enhanced indirectly as shown
in Chesapeake Bay through the creation of anoxic waters and oxidation of hydrogen
sulfide (Cai et al.. 2017b).
Ocean acidification, which can be exacerbated by elevated N input, is projected to alter
marine habitat, have a wide range of effects at the population and community level and
affect food web processes (Mostofa et al.. 2016; Andersson et al.. 2015; Gavlord et al..
2015; Sunda and Cai. 2012; Donev et al.. 2009). A major finding of The West Coast
Ocean Acidification and Hypoxia Science Panel was that while the dominant cause of
ocean acidification in the region is global CO2 emissions, local discharge of nutrients and
organic C is a factor exacerbating ocean acidification and hypoxia (Chan et al.. 2016).
Nutrient additions can trigger algal blooms, which upon decomposition, further reduce
oxygen levels in the water column and thus also lower the pH.
10-54
DRAFT: Do Not Cite or Quote

-------
Increasing
atmospheric
CO,
Loading
Algal blooms
from nutrient
inputs ,
Adverse
biological
impacts
/ Increasing \
[COz]/lower pH
as regulated by
^temperature and
salinity S
Microbial \
respiration of
organic
v matter /
Lower 02/
hypoxia
C02 = carbon dioxide; N = nitrogen; 02 = oxygen.
Modified from: Sunda and Cai (2012).
Figure 10-10 Pathway of nutrient-enhanced coastal acidification from nitrogen
loading to biological effects. Both microbial respiration of organic
matter and increasing atmospheric carbon dioxide lower pH of
coastal waters.
1	Acidification of coastal waters may cause varying degrees of harm to marine organisms
2	that produce calcium carbonate shells or skeletons, including oysters, clams, sea urchins,
3	shallow water corals, and calcareous plankton (Mostofa et al.. 2016; Gledhill et al.. 2015;
4	Pfister et al... 2014; Kroekcr et al.. 2013; Sunda and Cai. 2012). The acidifying process
5	decreases the saturation state of the two minerals (aragonite and calcite) that most
6	bivalves use to form their shells (Barton et al.. 2015; Barton et al.. 2012). Lower pH of
7	seawater can also potentially impair physiological energetics and spawning capacity of
8	shellfish (Xu et al.. 2016b). Commercially important species from the New England
9	coastal region have shown biological responses to changes in the carbonate system
10	associated with ocean and coastal acidification (Gledhill et al.. 2015). Already, there are
June 2018
10-55
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
declines in oyster production on the U.S. West Coast due to the inability of oysters to
create shells due to acidification (Chan et al.. 2016; Barton et al.. 2015; Hettinger ct al..
2012). However, other factors can confound the response of these sensitive organisms.
More research is needed to accurately determine the interactions and combined impact of
ocean acidification and N enrichment and other stressors on U.S. ecosystems like coral
reefs and estuaries with sensitive shellfish populations (Kroekeretal.. 2013).
In a recent modeling study of risk of food web and fisheries to ocean acidification in the
California current nearshore, invertebrates and associated fishery revenues are expected
to be impacted to a greater extent than pelagic species (Marshall et al.. 2017). Although
this modeling study did not take N inputs or other measures into account, invertebrates
that live in or on the benthic layer (crabs, shrimps, bivalves) will experience the strongest
effects of ocean acidification, followed by those species that consume benthic
invertebrates such as demersal fish and Dungeness crabs (Metacarcinus magister).
However, the model also showed that different species of groundfish reacted differently
to the changing pH, so some species and associated fishery dynamics may be impacted to
a greater degree than others.
Research on costressors associated with conditions of coastal acidification and
eutrophication suggest that interactions between elevated CO2, decreasing pH, and
nutrient inputs are complex. In macroalage Ulva pertusa grown under conditions of low
pH and high NH44", both growth rate and NH4+ uptake were significantly higher than in
treatments with higher pH and lower nutrient enrichment (Kang and Chung. 2017).
Young and Gobler (2016) tested the effects of elevated concentrations of CO2 alone, and
in combination with elevated nutrient (N + P) levels, on the growth rates of two common
species of opportunistic macroalgae, the rhodophyte Gracilaria and the chlorophyte
Ulva. Results showed that higher levels of pCOi significantly enhanced the growth rates
of both types of macroalgae, and the combination of enriched N + P and pCOi did appear
capable of synergistically promoting the growth of Ulva. Given that eutrophication can
yield elevated levels of pCOi. this study suggests that the overgrowth of macroalgae in
eutrophic estuaries can be directly promoted by acidification.
Eelgrass productivity and growth are predicted to increase under conditions of elevated
CO2 because photosynthesis in these plants is thought to be currently limited by CO2
(Koch et al.. 2013; Alexandre et al.. 2012; Palacios and Zimmerman. 2007). However,
seagrass productivity responses to combined stressors ofpC02 and NO3 were found to
be variable between different species (Ow et al.. 2016). In mesocosm experiments with
the seagrass Zostera noltii, photosynthesis increased under enriched CO2, but NO3 and
NH4+ uptake and growth rate were not significantly affected, indicating that seagrass
10-56
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
production may be limited by N availability under increased CO2 (Alexandre ct al..
2012V
Combined effects of NO3 enrichment, acidification, and temperature resulted in the
highest reproductive success for the green tide-forming algae Ulva rigida, pointing to
potentially more severe green tides under future scenarios than under present conditions,
especially in areas where eutrophication is a concern (Gao et al.. 2017). A decline in
calcification of the coral Marginopora rossi was observed at a pH of 7.6 alone or in
combination with nutrients (Rcvmond et al.. 2013). In NO3 enriched seawater, the
number of zooxanthellae cells in the host coral increased but with no observable benefits
on coral growth. A threshold for ocean acidification at pH 7.6 alone or in combination
with eutrophication will lead to a decline inM rossi calcification.
Models show that while the impact of each acidification pathway (N enrichment and
atmospheric CO2 dissolution) may be moderate, the combined effect of the two may be
much larger than would be expected by just adding the effects of each pathway together
(Sunda and Cai. 2012; Cai et al.. 2011c). These models, which incorporate the expected
future higher atmospheric CO2 levels and current levels of eutrophication, show dramatic
increases in acidification of coastal waters below the pycnocline. Model predictions
agreed well with pH data from hypoxic zones in the northern Gulf of Mexico and both
the modeled and the measured decreases in pH are well within the range shown to impact
marine fauna (Sunda and Cai. 2012). Data from the northern Gulf of Mexico revealed a
significant positive correlation between subsurface water pH and O2 concentration,
linking acidification to low O2 levels from organic matter decomposition (Cai et al..
2011c). Results from a coupled physical-biogeochemical model in the Gulf of Mexico
point to eutrophication driven by river nutrient input as an important factor in
acidification and hypoxia generation (Laurent et al.. 2017). Findings from the model
showed that acidified waters are predicted in a thin layer close to the bottom of the Gulf,
similar to the vertical distribution of the seasonal hypoxia zone.
The increase in atmospheric CO2 dissolution is also expected to alter the N cycle in the
ocean, with implications for the entire ecosystem. Acidification will result in decreased
rates of nitrification, which combined with expected increasing N deposition, is expected
to cause the NH44" concentration of the water to rise (Lefebvre et al.. 2012). A study of the
coccolithophore Emiliania huxleyi, which plays a major role in the global carbon cycle by
regulating the exchange of CO2 across the ocean-atmosphere interface, found that higher
levels ofNH4+ can affect the morphology and calcification of the coccolithophore. The
combined effect of higher NH4+ levels and greater acidification could reduce calcification
rates, suggesting a greater alteration to the ocean's carbon cycle due to the combined
effect of increased NH4 /NO3 and acidification (Lefebvre et al.. 2012).
10-57
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
10.6.
Summary of Thresholds and Levels of Deposition at Which
Effects Are Manifested
Studies linking changes in estuary nutrient status to atmospheric deposition have been
limited, although some states are addressing atmospheric inputs as part of their
development of Total Maximum Daily Loads (TMDLs) (Linker et al.. 2013). Since the
2008 ISA, additional thresholds of response to N have been identified for biological
indicators. In general, the identification and application of indicators and thresholds to
assess eutrophication in coastal areas is more highly developed in Europe than in the U.S.
due to the implementation of the EU's Water Framework Directive (Zaldivar et al..
2008). Here we limit discussion to indicators and thresholds that have been applied to
U.S. water bodies.
For chlorophyll a, both NEEA ASSETS and U.S. EPA NCCA have a similar range for
categorizing eutrophic conditions; however, ASSETS uses the 90th percentile chlorophyll
concentration of annual data and U.S. EPA NCCA uses growing season (June-October)
values [(Borjaetal.. 2012); Table 10-21. For ASSETS, chlorophyll concentrations of
0-5 |ig/L is a water body with low risk of eutrophication, 5-20 |ig/L is moderate,
>20 |ig/L is high with >60 |ig/L indicating hypereutrophic conditions. NCCA identified
chlorophyll concentration of >20 |ig/L as an indicator of an estuary in poor condition.
Chlorophyll concentrations between 5-20 |ig/L are classified as fair and chlorophyll
concentration 0-5 |ig/L was good for sites located in the Northeast, Southeast, Gulf, West
Coast, and Alaska (U.S. EPA. 2012b). Values were lower for sites in Hawaii, Puerto
Rico, U.S. Virgin Islands, American Samoa, and Florida Bay (0.5 (ig/L good, 0.5-1 (ig/L
fair, >1 |ig/L. poor).
For DO, ASSETS uses the 10th percentile of annual data. DO concentrations of 0 mg/L
are anoxic, 0-2 mg/L are indicative of hypoxic conditions, and 2-5 mg/L are biologically
stressful conditions (Devlin et al.. 2011; Bricker et al.. 2007). Spatial coverage (range of
0-10% [low] to >50% [high]) and frequency of occurrence (persistent, periodic,
episodic) are also included in determining reference thresholds for DO with ASSETS.
For U.S. EPA NCCA, the cutpoint used for poor DO condition is <2 mg/L in bottom
waters (U.S. EPA. 2016g. 2012b). Because many states use higher concentrations, the
NCCA considers concentrations between 2 and 5 mg/L as fair, and >5 mg/L as good.
Under the U.S. Clean Water Act, states are in the process of developing numeric nutrient
criteria to better define levels of N and P that affect estuaries and coastal marine waters
(Appendix 7.2.10). The numeric values include both causative (N and P) and response
(chlorophyll a, biocriteria) variables to assess eutrophic conditions. For progress toward
state development of numeric criteria for N see http://cfpub.epa.gov/wqsits/nnc-
June 2018
10-58
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
development/. Atmospheric deposition as a source of N may represent a potential target
for remediation such as TMDL, nutrient budgets, and allocations.
The Chesapeake Bay 2010 TMDL load reductions include atmospheric deposition. This
is the first time atmospheric N loads to tidal waters have been included in a TMDL for
reduction (U.S. EPA. 2010). For Chesapeake Bay, a numerical chlorophyll criteria
ranging from 1.4 to 15 mg/m3 (|ig/L) based on seasonal mean across salinity zones was
proposed as a restoration goal with 90th percentile compliance limits ranging from 4.3 to
45 mg/m3 [fig/L; (Harding et al.. 2014)1. These values were based on relationships of
chlorophyll a concentrations to DO, water clarity, and presence of toxic algae.
Chlorophyll a concentration ranges of 7.2 to 11 mg/m3 (|ig/L: May-August mean) in the
bay and 9 to 14 mg/m3 (|ig/L: annual mean) in the tributaries would be protective of low
DO conditions. For SAV growth, a mean chlorophyll a concentration of 7.9-12 mg/m3
(fig/L) with a threshold of 19-28 mg/m3 (|ig/L) during the growing season would allow
for sufficient light conditions. For decreased risk for toxic cyanobacteria, a mean summer
chlorophyll a concentration of 15 mg/m3 and a threshold of 25 mg/m3 (|ig/L) was
recommended.
For seagrasses in New England estuaries, a threshold of N loading (N export of
wastewater, fertilizer, and atmospheric deposition from the watershed combined with
direct atmospheric deposition to waters) >100 kg N/ha/yr was identified. Eelgrass is
essentially absent in areas with these levels of N loading, and levels above 50 kg N/ha/yr
are likely to impact habitat extent ITable 10-4; (Latimer and Rego. 2010)1. These values
were based on literature threshold values from Bowen and Valiela (2001). Hauxwell et al.
(2003). Li et al. (2007). and Latimer and Rego (2010). Greaver et al. (2011) identified the
range of 50-100 kg/N/ha/yr total N loading as the empirical critical load for loss of
eelgrass based on (Latimer and Rego. 2010). These threshold levels for seagrasses were
developed in shallow, poorly flushed systems that may not be universally representative
of all estuaries.
10.7. Summary and Causal Determinations
Causal determinations between N loading in coastal areas and biological effects of N
enrichment (Appendix 10.7.1) and nutrient-enhanced coastal acidification
(Appendix 10.7.2) are presented in this summary along with supporting evidence for
these relationships.
June 2018
10-59
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
10.7.1.
Nitrogen Enrichment
In the 2008 ISA, the body of evidence was sufficient to infer a causal relationship
between N deposition and the alteration of species richness, species composition, and
biodiversity in estuarine systems (U.S. EPA. 2008a'). The strongest evidence for a causal
relationship was from changes in biological indicators of nutrient enrichment
(chlorophyll a, macroalgal abundance, HABs, DO, SAV) and N was recognized as the
major cause of harm to the majority of estuaries in the U.S. (Brickcr et al.. 2008; NRC.
2000). Phytoplankton are the base of the coastal food web, and increases in primary
producer biomass and altered community composition associated with increased N can
lead to a cascade of direct and indirect effects at higher trophic levels. For this ISA, new
information is consistent with the 2008 ISA, and the causal determination has been
updated to reflective more specific categories of effects. The body of evidence is
sufficient to infer a causal relationship between N deposition and changes in biota,
including total primary production, altered growth, total algal community biomass,
species richness, community composition, and biodiversity due to N enrichment in
estuarine environments.
Since the 2008 ISA, additional evidence has shown that reduced forms of atmospheric N
play an increasingly important role in estuarine and coastal eutrophication and HAB
dynamics. In many parts of the U.S., especially the Southeast, Midwest and Mid-Atlantic
deposition of reduced N has increased relative to oxidized N in last few decades. The
form of N delivered to some coastal areas of the U.S. is shifting from primarily NOs to
an increase in reduced forms of N. The increase in highly bioreactive reduced N from
deposition and other sources is often a preferred form of N for specific phytoplankton
functional groups (e.g., cyanobacteria, dinoflagellates) including harmful species.
(Glibert et al.. 2016; Paerl et al.. 2000). Atmospheric inputs to estuaries are
heterogeneous across the U.S., ranging from <10 to approximately 70% of the N inputs
(Table 7-9). so deposition may play a significant role in altering nutrient dynamics in
some coastal systems (i.e., Mid-Atlantic, Southeast) that support an increasing human
population and spatial expansion of industrial scale animal operations (Li et al.. 2016d).
Chlorophyll a is a broadly applied indicator of phytoplankton biomass and used as a
proxy for assessing effects of estuarine nutrient enrichment. It can signal an early stage of
water quality degradation related to nutrient loading and is incorporated into water
quality monitoring programs and national-scale assessments, including U.S. EPA's
NCCA and the NEEA. In general, 0-5 |ig/L chlorophyll concentration is considered to be
good, chlorophyll concentrations between 5-20 |ig/L are classified as fair, and >20 |ig/L
indicates an estuary in poor condition. Phytoplankton sampling and sediment core
analysis have shown changes in phytoplankton community structure in estuaries with
June 2018
10-60
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
elevated N inputs. These shifts at the base of the food web to species that are not as
readily grazed (e.g., cyanobacteria, dinoflagellates) have a cascade of effects that include
poor trophic transfer and an increase in unconsumed algal biomass, which could
stimulate decomposition, O2 consumption, and the potential for hypoxia (Paerl ct al..
2003). Although availability of N is an important factor for productivity in estuaries,
many newer studies have emphasized the role of physical and hydrologic factors in
increase of chlorophyll a and other indicators of N enrichment (Hart et al.. 2015; Turner
et al.. 2015; Wilkerson et al.. 2015; Glibert et al.. 2014; Kennison and Fong. 2014;
Rothenberger et al.. 2014; Paerl et al.. 2010; Yang et al.. 2008).
Incidence of HAB outbreaks continues to increase in both freshwater and coastal areas, a
problem that has been recognized for several decades (U.S. EPA. 2008a; Bricker et al..
2007; Paerl et al.. 2002; Paerl and Whitall. 1999; Paerl. 1997). Of the 81 estuary systems
for which data were available in the NEEA reported in the 2008 ISA, 26 exhibited a
moderate or high symptom expression for nuisance or toxic algae (Bricker et al.. 2007).
HAB bloom conditions and effects of HAB toxins on coastal biota have been further
characterized since the 2008 ISA (Wood et al.. 2014; Miller et al.. 2010). The form of N
affects phytoplankton growth and toxin production of some HAB species. CyanoHAB
presence has been documented in estuaries along the U.S. Mid-Atlantic and Southeast
coasts as well as Gulf coast estuaries (Preece et al.. 2017).
In addition to phytoplankton, seaweed growth is also stimulated by increased N inputs.
Macroalgal blooms can smother benthic organisms and corals and contribute to the loss
of important SAV by blocking the penetration of sunlight into the water column. Studies
published since the 2008 ISA provide further evidence that macroalgae respond to the
form of N, with some species showing greater assimilation and growth rates with NH44"
than with NO3 (Wang et al.. 2014a; Ale etal.. 2011).
Since the 2008 ISA, seagrass coverage is improving or stable in some estuaries like
Tampa Bay and Chesapeake Bay, while many areas continue to see declines in seagrass
extent. Loss of SAV habitat can lead to a cascade of ecological effects because many
organisms are dependent upon seagrasses for cover, breeding, and as nursery grounds.
SAV is often at a competitive disadvantage under N enriched conditions due to the fast
growth of opportunistic macroalgae that preferentially take up NH4+ and can block light
from seagrass beds. Increased epiphyte loads on the surface of macrophytes from nutrient
enrichment may reduce growth of SAV. Since the 2008 ISA, additional studies are
available on the relationship between N loading and SAV abundance (Bovnton et al..
2014; Orth et al.. 2010; Ruhl and Rvbicki. 2010). and several studies have identified
specific regional thresholds lYBenson et al.. 2013; Latimer and Rego. 2010); Table 10-41.
10-61
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Greaver et al. (2011) identified the range of 50-100 kg/N/ha/yr total N loading as the
empirical critical load for loss of eelgrass based on (Latimer and Rego. 2010).
Increased algal biomass associated with nutrient over-enrichment leads to increased
decomposition of organic matter, which decreases DO. Oxygen depletion largely occurs
only in bottom waters under stratified conditions, not throughout the entire water column.
A variety of ecological impacts are associated with low DO (Figure 10-4). and this
indicator has been used in national assessments of coastal condition including U.S. EPA's
NCCA and the NEEA. DO concentrations of 0 mg/L are anoxic, 0-2 mg/L are indicative
of hypoxic conditions, and 2-5 mg/L are biologically stressful conditions (Devlin et al..
2011; Bricker et al.. 2007). For example, many fishes are absent at DO levels below
2 mg/L, and mortality of tolerant organisms can occur at approximately 0.5 mg/L.
Hypoxia has recently been shown to affect reproductive parameters in fish which may
lead to effects at the population level. In laboratory conditions, increased turbidity
associated with eutrophic conditions has been shown to alter fish reproductive behaviors
(Candolin. 2009). Macroinvertebrate community structure is also affected by the duration
and severity of hypoxia. A shift to benthic organisms with shorter life spans and smaller
body size has been observed in coastal areas that experience severe seasonal hypoxia
(Diaz and Rosenberg. 2008). Reduced species density and diversity in the northern Gulf
of Mexico are linked to the duration of hypoxic events (Baustian and Rabalais. 2009).
Post-2007 literature includes additional information on the extent and severity of
eutrophication and hypoxia in sensitive regions. Diaz et al. (2013) assessed global
patterns in hypoxia from the 1960s to 2011. Areas of eutrophication-related hypoxia are
found on the U.S. East and West Coasts and the Gulf of Mexico (Figure 10-5). Overall,
the extent of hypoxia is increasing globally with some areas showing signs of
improvement. In the 2008 ISA, the largest zone of hypoxic coastal water in the U.S. was
the northern Gulf of Mexico on the Louisiana-Texas continental shelf (U.S. EPA. 2008a).
This area, which forms annually between May and September, continues to be the largest
in the U.S. and the second largest in the world, averaging about 16,500 km2 (10,250 mi2)
in size (Dale et al.. 2010; Jewett et al.. 2010). In the summer of 2017, the hypoxic zone in
the Gulf was the largest ever measured at 14,123 km2 (8,776 mi2) (U.S. EPA. 2017e).
Atmospheric deposition from watersheds in the Mississippi/Atchafalaya River basins
contributes approximately 16 to 26% of the total N load in the Gulf of Mexico
(Robertson and Saad. 2013; Alexander et al.. 2008). Long Island Sound also experiences
periods of anoxia in some years. In other U.S. coastal systems, the incidence of hypoxia
is increasing; however, the severity of DO impacts are relatively limited temporally and
spatially (Jewett et al.. 2010; U.S. EPA. 2008a; Bricker et al.. 2007). In the Pacific
Northwest, coastal upwelling can be a large source of nutrient loads, and advection of
10-62
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
upwelled water can introduce into estuaries hypoxic water that is not related to
anthropogenic sources.
The indicators described above (chlorophyll a, HABs, macroalgal abundance, DO, SAV,
and benthic diversity) have been incorporated into indices that describe eutrophic
conditions in coastal areas. In the 2008 ISA, the ASSETS ECI developed for the NEEA
was used as an assessment framework for eutrophic conditions in coastal U.S. estuaries
(Bricker et al.. 2007). The NCCA incorporates indicators that include chlorophyll a and
DO to assess U.S. waters (U.S. EPA. 2016g. 2012b). Additional indices of overall
condition of estuarine functioning that incorporate biological indicators of eutrophication
have since been developed both in the U.S. and other countries (Boria et al.. 2012; Devlin
et al.. 2011; Boria et al.. 2008). Comparisons of these frameworks have led to the
identification of more robust and representative methods to measure estuarine response,
including incorporation of annual data, frequency of occurrence, spatial coverage,
secondary biological indicators, and a multicategory rating scale (Devlin etal.. 2011).
In addition to having effects on biota in estuarine environments, N enrichment is one of
several factors linked to increased disease susceptibility, bleaching, and reduced
calcification rate in corals (Appendix 10.4.2). Near-coastal coral reefs in the U.S. occur
off southern Florida, Texas, Hawaii and U.S. territories in the Caribbean and Pacific. The
threatened status of staghorn coral (Acropora cervicornis) and elkhorn coral (Acropora
palmata) under the U.S. Endangered Species Act has been linked to indirect N pollution
effects, specifically low DO and algal blooms that alter habitat, and to other non-nutrient
stressors (Hernandez et al.. 2016).
10.7.2. Nutrient-Enhanced Coastal Acidification
Since the 2008 ISA, N enrichment has been recognized as a possible contributing factor
to increasing acidification of marine environments. Dissolution of atmospheric
anthropogenic CO2 into the ocean has led to long-term decreases in pH (Wallace et al..
2014; Orr et al.. 2005). With increasing N inputs to coastal waters, CO2 in the water
column is produced from degradation of excess organic matter from changing land use,
as well as respiration of living algae and seagrasses, which in turn can make the water
more acidic (Wilson et al.. 2016; Howarth et al.. 2014; Wallace et al.. 2014; Wetz and
Yoskowitz. 2013; Sunda and Cai. 2012; Cai et al.. 2011c; Howarth et al.. 2011).
Nutrient-enhanced coastal acidification has been documented in systems with strong
thermal stratification with spatial or temporal decoupling of production and respiration
processes. Models show that while the impact of each acidification pathway (N
enrichment and atmospheric CO2 dissolution) may be moderate, the combined effect of
June 2018
10-63
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
the two may be much larger than would be expected by just adding the effects of each
pathway together (Sunda and Cai. 2012; Cai etal.. 2011c). The body of evidence is
suggestive of, but not sufficient to infer a causal relationship between N deposition
and changes in biota, including altered physiology, species richness, community
composition, and biodiversity due to nutrient enhanced coastal acidification.
Ocean acidification, which can be exacerbated by elevated N input, is projected to alter
marine habitat, have a wide range of effects at the population and community level and
affect food web processes (Mostofa et al.. 2016; Andersson et al.. 2015; Gavlord et al..
2015; Sunda and Cai. 2012; Donev et al.. 2009). Organisms that produce calcium
carbonate shells, such as calcareous plankton, oysters, clams, sea urchins, and coral are
impacted by increasing acidification of waters (Barton et al.. 2015; Kroeker et al.. 2013;
Barton et al.. 2012; Hettinger et al.. 2012). Decreased saturation rates of aragonite and
calcite, two minerals needed in shell formation, are observed in acidic conditions. These
physiological responses to acidifying conditions in coastal waters may lead to effects at
the population level. Changes in the carbonate system, including decreased pH, have
been shown to elicit biological responses in commercially important species from the
New England coast and there are documented declines in oyster production on the U.S.
West Coast linked to ocean acidification. Research on costressors associated with
conditions of coastal acidification and eutrophication suggest that interactions between
elevated CO2, decreasing pH, and nutrient inputs are complex.
10-64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
APPENDIX 11. NITROGEN ENRICHMENT EFFECTS
IN WETLANDS
This appendix describes the effects of N deposition on wetland ecosystems. The
introduction contains an overview of the relationship between N deposition and wetland
endpoints, as well as the wetland classification system used in this appendix
(Appendix 11.1). Regional sensitivity in wetlands is related to position within the
watershed (Appendix 11.2). N deposition causes changes to biogeochemical pools and
processes in wetlands, specifically to N (Appendix 11.3.1) and C cycling
(Appendix 11.3.2). Nitrogen eutrophication affects wetland primary producers via
alteration of aboveground plant biomass (Appendix 11.4). plant stoichiometry and
physiology (Appendix 11.5). plant architecture (Appendix 11.6). and plant demography
(Appendix 11.7). Nitrogen eutrophication also alters wetland biodiversity via changes to
plant community composition (Appendix 11.8.1) and to consumer communities
(Appendix 11.8.2. There are a number of critical loads published for wetland ecosystems
(Appendix 11.9).
11.1. Introduction
The 1993 Oxides of Nitrogen Air Quality Criteria document (hereafter referred to as the
1993 AQCD) and the 2008 ISA for Oxides of Nitrogen and Sulfur—Ecological Criteria
(hereafter referred to as the 2008 ISA) evaluated the effects of nitrogen deposition on
wetland ecosystems. The 1993 AQCD found that the three main ecological effects of N
deposition on wetland ecosystems were reduced biodiversity, modified microbial
processes, and increased primary production. The 2008 ISA supported and extended the
conclusions in the 1993 AQCD, especially with regard to the effects of N deposition on
species diversity, and also found evidence for alterations of ecosystem nitrogen and
carbon cycling. The 2008 ISA found that the evidence was sufficient to infer a causal
relationship between N deposition to wetland ecosystems and the alteration of
biogeochemical cycling, as well as the alteration of species richness, species composition,
and biodiversity. New evidence published between 2008-2015 from observational
studies, experimental N additions in the field and in mesocosms, and reanalysis of large
data sets supports and extends the conclusions of the 2008 ISA. The body of evidence is
sufficient to infer a causal relationship between N deposition and the alteration of
biogeochemical cycling in wetlands. In addition to new information on species
biodiversity, there is recently published evidence of N deposition effects on endpoints not
covered in the 2008 ISA, including alterations to plant physiology and plant architecture.
June 2018
11-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
The body of evidence is sufficient to infer a causal relationship between N deposition
and the alteration of growth and productivity, species physiology, species richness,
community composition, and biodiversity in wetlands.
Wetland vegetative communities are adapted to high levels of natural organic acidity, so
S and N deposition are unlikely to cause any acidification-related effects at levels of
acidic deposition commonly found in the U.S. (U.S. EPA. 2008a. in Annex B).
Sensitivity of wetlands to N deposition as a nutrient are well documented in the 2008 ISA
(Appendix 11.2). Hydrologic pathways are often the same pathways of N input;
therefore, they are useful for discussing the N sources and sensitivity to atmospheric N
deposition. In general, wetland types (Table 11-1) that receive most of their total N input
from the atmosphere are most sensitive to atmospheric deposition. Nearly all new N
comes from atmospheric deposition in ombrotrophic bogs because these wetlands only
receive water inputs via precipitation. They develop where precipitation exceeds
evapotranspiration and where there is some impediment to drainage of the surplus water
(Mitsch and Gosselink. 1986). Freshwater fens, marshes, and swamps are characterized
by ground and surface water inputs that are often on the same order of magnitude as
precipitation (Koerselman. 1989). Lastly, estuarine and coastal wetlands receive water
from precipitation, ground/surface water, and marine/estuarine sources. Therefore, bogs
(and presumably vernal pools) are among the most vulnerable wetlands to the
nutrient-enrichment effects of N deposition (Krupa. 2003).
Table 11-1 Wetland classification used in the Integrated Science Assessment.
Wetland Term Definition (adapted from, Wakelev. 2002: Mitsch and Gosselink. 2000: Cowardin et al„ 1979)
Soil-based classification
Peat	Substrate consisting of partially decomposed plant litter. In bryophyte-dominated wetlands (bogs
and fens), this layer is composed of dormant or dead moss. In marshes and swamps, peat consists
of dead litter, including plant leaves, stems, and roots, as well as undifferentiated soil organic
matter and sediments.
Peatland
Any wetland that accumulates or has historically accumulated organic C stores in the substrate.

Net primary productivity and accumulation of fixed C exceed decomposition rates. In Europe, the

synonymous term "mire" is used.
Hydrology-based classification
Permanent
Wetland where a zone of the soil profile or peat profile remains saturated with water. The deeper
wetland
soil profile is typically saturated, anoxic, and often has a high percentage of organic matter.
Intermittent
Wetland where soils are saturated with water for short periods of time, annually or every few years;
wetland
soils tend to have higher mineral contents than do permanent wetland soils.
June 2018
11-2
DRAFT: Do Not Cite or Quote

-------
Table 11-1 (Continued): Wetland classification used in the Integrated Science
Assessment.
Wetland Term Definition (adapted from, Wakelev. 2002: Mitsch and Gosselink. 2000: Cowardin et al.. 1979)
Intertidal	Wetlands where soils or sediments are inundated with water at high tide, and are exposed to air at
wetland	low tide. This includes salt marshes, mangroves, and unvegetated mudflats.
Soil-, hydrology-, and vegetation-based classification
Bog or peat Defined by vegetation. A wetland with high moss cover and a depth profile with high organic
bog	content (peat), sometimes comprised of compressed senesced mosses. Ericaceous shrubs and
certain graminoids are also particularly adapted for bog conditions.
Ombrotrophic A peatland dominated by acidophilic mosses, particularly Sphagnum mosses. Water is derived
or oligotrophic entirely from precipitation, with high [DOC], low pH, and low nutrient concentrations.
bog
Mesotrophic, A peatland that supports herbaceous and woody plant species with a wetland surface composed of
poor, or	moss species. Water sources include precipitation, surface water, and groundwater. Water pH
intermediate	(acidic to circumneutral) and nutrient concentrations are intermediate between ombrotrophic bog
fen	and minerotrophic fen.
Minerotrophic,	A peatland dominated by herbaceous graminoid species typical of marshes, with a mat composed
rich, eutrophic,	of moss species. Water sources include precipitation, surface water, and groundwater. Water has
or calcareous	higher pH (circumneutral to alkaline), as well as higher concentrations of calcium ions and other
fen	nutrients, than other bog types.
Swamp	Defined by vegetation. A wetland dominated by woody plants, particularly tree and shrub species
with physiological adaptations to occasional or constant soil inundation.
Mangrove An intertidal swamp. A subtropical or tropical swamp dominated by halophytic trees and shrubs.
Tidal inundation by ocean or estuarine waters is frequent. In the U.S., dominant species are trees
in the Avicennia and Rhizophora genera. Mangroves are important nursery habitats for marine
animal species.
Marsh	Defined by vegetation. A wetland dominated by herbaceous plants with physiological adaptations
to occasional or constant soil inundation.
Freshwater Marsh dominated by herbaceous plant community composed of species with physiological
marsh	adaptations to soil inundation. Includes marshes in lakes, rivers, and nonsaline regions of
estuaries. Salt- and sulfide-intolerant marsh plant species are common. Soils may be peat or
mineral sediments.
Pothole	A type of freshwater marsh. A shallow pond with herbaceous and submerged aquatic vegetation,
common in the prairie ecosystems of the Dakotas and central Canadian plains. A pothole is often
not connected by surface water flow to its watershed; water sources include precipitation and
groundwater.
Tidal marsh An estuarine marsh. Used in this document to describe estuarine marshes in which water level
fluctuates in response to ocean tides but marsh water has low or no salinity (i.e., water sources are
freshwater). Depending on physical location in the estuary, occasional saltwater intrusion may alter
biogeochemistry and plant community. A mix of saline-tolerant and saline-intolerant plant species
is common.
Salt or coastal Halophytic herbaceous species are dominant. Tidal inundation by ocean water structures the plant
marsh	community, which typically forms zones of single species monocultures or low-diversity
communities of plants with similar tolerances for salt and water stress. Plant roots trap alluvial
sediments to build and maintain the elevation of the marsh surface, or platform. Salt marshes are
important nursery habitats for marine animal species.
June 2018	11-3	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
Table 11-1 (Continued): Wetland classification used in the Integrated Science
Assessment.
Wetland Term Definition (adapted from, Wakelev. 2002: Mitsch and Gosselink. 2000: Cowardin et al.. 1979)
Intermittent Defined by hydrology and vegetation. Wetland where soils are saturated, with soil processes
typical of aquatic sediments, for short periods annually or every few years; inundation occurs with
sufficient frequency to prevent persistence of flooding-intolerant plant species.
Vernal pool An area with mineral soils which is flooded by rainwater or rising water tables and then isolated
from surface water flow as it slowly dries up. These seasonally inundated areas are important
nursery habitat for a number of invertebrate and vertebrate species.
Riparian	An area with mineral soils, and a high water table due to close proximity to a river, stream, or lake,
wetland	Periodic inundation occurs when water levels rise. Vegetation is a mix of terrestrial plants and
plants with physiological adaptations to occasional soil inundation.
C = carbon; DOC = dissolved organic carbon.
11.2. Regional Sensitivity
There is no new information on regional sensitivity of wetlands to N deposition, and
models of sources of N to wetland ecosystems are not yet available. In the 2008 ISA,
wetlands were described in order of sensitivity to N deposition. Bogs and fens are among
the most sensitive wetland ecosystems to N deposition. In the U.S., peat-forming bogs
and fens are most common in glaciated areas, especially in portions of the Northeast and
Upper Midwest (U.S. EPA. 1993). N deposition was projected to drastically change
species composition based on experimental results in European fens (Pauli et al.. 2002;
Aerts and de Caluwe. 1999). N input and output rates of freshwater or riparian marshes
and swamps are intermediate between bogs and coastal marshes. N deposition increased
primary productivity 30% and increased methane emissions in freshwater marshes.
Atmospheric N inputs contribute to eutrophication problems in coastal marshes at many
locations through direct deposition to the marsh, indirect deposition to the watershed, or
direct deposition to estuaries or coastal waters followed by tidal delivery of aqueous N
loads to wetlands (see Table 7-9 for estimates of N deposition to estuaries). However,
marine inputs of N are typically higher than direct atmospheric input to coastal wetlands.
The National Wetlands Condition Assessment (NWCA) was conducted by U.S. EPA in
2011 to characterize the condition of North American wetlands under multiple
anthropogenic stressors and to provide a baseline for future assessments. The final report
did not provide soil N or surface water ammonia, nitrate-nitrite, or total N, although these
data were collected as part of the survey and are available (U.S. EPA. 2016i). Ecological
endpoints that the NWCA identified as correlated with total N loading are the Normative
Plant Stressor Indicator (NPSI; see Appendix 11.8.1.3 for an explanation of the link
June 2018
11-4
DRAFT: Do Not Cite or Quote

-------
1	between N loading and non-native plants) and microscystin concentrations (see
2	Appendix 9). although the dataset includes a number of other metrics that may respond to
3	N deposition. Nationally, 39% of wetland area was estimated to experience moderate to
4	very high stress based on the NPSI, with stresses particularly high in wetlands of the
5	Interior Plains and West (see Figure 11-1). Microcystis a toxin produced by
6	cyanobacteria in response to available N, was detected in low concentrations in surface
7	water at 12% of national wetland sites, with particularly high detection rates (34%) in the
8	Interior Plains (U.S. EPA. 2016i).
National
Coastal
Plains
Eastern Mtn. &
Upper Midw.
hfc
CJV
&SK
'\r \>

Interior
Plains
West
Nonnative Plants
Percent Area
Nonnative Plants
Area
20 40 60 80 100 0
Percent Area
i i Lnwi I Moderate
High I
20,000,000
Area
Very High
40,000,000
Source: U.S. EPA (2016iV
Figure 11-1 Estimated extent of wetlands stressed by nonnative plants as
determined by Nonnative Plant Stressor Indicator, at a national or
regional level.
June 2018
11-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
11.2.1. Climate Modification of Ecosystem Response to Nitrogen
Changes in mean annual temperature and frequency and magnitude of precipitation will
affect the responses of all wetlands to N loading; changes in sea level will affect the
response of salt marsh, mangroves, and freshwater tidal wetlands. Temperature effects on
wetlands have been demonstrated in European bog ecosystems, where increased
temperatures increased cover of woody species and decreased Sphagnum moss cover just
as N deposition does (Hcdw all et al.. 2017). In metaregressions of Sphagnum moss
response to N addition and temperature, the two factors synergistically depressed
Sphagnum production, with an 1°C increase in summer temperature having an impact on
Sphagnum equivalent to an additional 40 kg N/ha/yr (Limpens et al.. 2011). Increasing
temperatures may strengthen N effects upon wetland ecosystems.
Hydrologic regimes are important controls on wetland cycling and productivity, so
changes in the magnitude and frequency of precipitation can have strong effects on
ecosystem N retention and C storage. The same metaregression of Sphagnum mosses
referenced above found that increased precipitation also increased Sphagnum sensitivity
to N addition-induced decreases in production (Limpcns et al.. 2011). Experimental
mesocosms modelling changes in precipitation to salt marshes found that precipitation
delivered in infrequent, heavy storm events decreased N retention and plant productivity,
even though storm events delivered a higher N deposition load to the marsh (Hanson et
al.. 2016; Oczkowski et al.. 2016). Shifts in precipitation towards less frequent, more
intense rain events may strengthen N deposition-induced decreases in salt marsh N
retention while weakening N deposition-induced increases in plant productivity.
Sea level rise will affect tidal wetlands through salt water intrusion into tidal freshwater
wetlands (Barcndrcgt and Swarth. 2013) and through increasing inundation of salt
marshes, both of which have the potential to change community composition and
decrease wetland extent. Experimental sea level rise of 10 cm increased N stimulation of
belowground and aboveground plant biomass in Kirkpatrick Marsh (see (Langlcv et al..
2013) in Appendix 11.3.2.1.2 and Appendix 11.4.1). but the response was
species-specific, with only one species responding to the N deposition. Sea level rise
effects on marsh response to N deposition are not well understood and may vary by
species.
11.3. Soil Biogeochemistry
The 2008 ISA described the relationship between N deposition and the alteration of
biogeochemical cycling in wetlands. N deposition alters N cycling by altering microbial
June 2018
11-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
communities, the rates at which they transform N between pools, ecosystem retention or
release of transformed N, and the emissions of the greenhouse gas nitrous oxide
(Appendix 11.3.1). N deposition alters C cycling by altering soil processes such as
decomposition and respiration (resulting in emissions of the greenhouse gases carbon
dioxide and methane), as well as by altering soil pools such as belowground biomass and
soil organic matter (Appendix 11.3.2). There is new evidence since 2008 of N deposition
effects upon biogeochemical cycling across wetlands, as well as in salt marshes,
mangroves, bogs, and riparian wetlands. The body of evidence is sufficient to infer a
causal relationship between N deposition and the alteration of biogeochemical
cycling in wetlands.
11.3.1. Nitrogen Pools and Processes
Water table depth tends to fluctuate more in wetlands than terrestrial ecosystems. The
variable depth creates dynamic anoxic-oxic boundaries and many varied niches for
different microbial communities. As a result, wetlands are important disproportionately to
their spatial area on the landscape (i.e., hotspots) for the microbial transformation of
nitrogen between oxidized and reduced forms. These nitrogen transformations can
improve water quality of downstream streams, lakes, and estuaries by removing N in the
soil or water of an ecosystem into the atmosphere. However, these transformations can
result in emissions of the potent greenhouse gas nitrous oxide. Alteration to N cycling or
N cycling microbial endpoints indicate changes to the important wetland function of
improving water quality.
N deposition contributes to total N load in wetlands. The chemical indicators that are
typically measured include NOs and NH4+ leaching, DON leaching, N mineralization,
and denitrification rates, including N2O emissions produced by incomplete
denitrification. N dynamics in wetland ecosystems are variable in time and among types
of wetlands and environmental factors, especially water availability (Howarth ct al..
1996a). A wetland can act as a source, sink, or transformer of atmospherically deposited
N (Devito et al.. 1989). and these functions can vary with season and hydrological
conditions. Vegetation type, physiography, local hydrology, and climate all play
significant roles in determining source/sink N dynamics in wetlands (Mitchell et al..
1996; Arhcimcr and Wittgrcn. 1994; Koerselman et al.. 1993; Devito et al.. 1989).
N mineralization (microbial transformation of organic N to inorganic forms of N) has
been shown to increase with N addition, and this can cause an increase in wetland N
export to adjacent surface water (Groffman. 1994). In general, ecosystem leaching losses
of NO3 from wetlands to downstream aquatic systems are small, although recent
June 2018
11-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
research suggests that N addition to peatlands increases export of N in both inorganic and
organic forms (Edokpa et al.. 2015). Bogs and fens in the Marcell Experimental Forest
that receive 2.8-4.7 kg N/ha/yr as atmospheric deposition export 0.26 (bogs) and 1.34 kg
N/ha/yr (fens), respectively (Hill et al.. 2016). Wetlands tend to have low N export to
surface water because anoxic zones within wetlands are favorable for microbial
denitrification of N from NOs to gaseous N forms. Elevated N inputs to wetlands will
often increase the rate of denitrification (Duffy and Kahara. 2011; Cooper. 1990;
Broderick et al.. 1988; Dierberg and Brezonik. 1983). because N additions to aquatic
environments with high organic C increase denitrifier abundance, activity, and
denitrification rates (Kim et al.. 2015b). This mitigates environmental effects associated
with increased N supply to soils and drainage waters; however, increased denitrification
may also increase the emissions of greenhouse gases (e.g., N2O) to the atmosphere.
Denitrification appears to be negligible in wetland environments that are typically
nutrient (including N) poor, such as some bogs and fens (Morris. 1991). Incomplete
denitrification and emission of N2O tend to be higher in freshwater tidal wetlands than in
saline wetlands like mangroves or salt marshes (Welti et al.. 2017). In salt marshes,
dissimilatory nitrate reduction to ammonium (DNRA) is an important process that cycles
N while retaining N within marsh sediments (see also Appendix 7.2.6.5). N addition
decreases DNRA while increasing nitrification and denitrification rates (Peng et al..
2016). decreasing N retention in salt marshes, and increasing N2O emissions (Chmura et
al.. 2016; Moseman-Valtierra et al.. 2011).
New studies summarized below in salt marsh, mangrove, peat bog, and riparian wetlands
have evaluated the effects of N loading/N addition on endpoints related to N cycling. In
addition, there is a new study on wetland N removal, synthesizing information across
wetland types.
11.3.1.1. Across Wetlands
N removal, defined as the sum of denitrification, plant uptake, and burial in sediments,
was evaluated by analyzing data from 109 wetlands distributed globally, including
agricultural wetlands, freshwater marshes, freshwater swamps, and coastal marshes
(Jordan et al.. 2011). Total N loads to these wetlands ranged from 0.2-90,480 kg N/ha/yr;
however, the sources of reactive N were not identified by the authors. Across this global
data set, wetland N removal as measured in outflow is proportional to wetland N load. N
removal efficiency was 26% higher in nontidal wetlands than in tidal wetlands.
The 2008 ISA found evidence that N deposition alters the emission of nitrous oxide
(N2O) and methane (CH4), which contribute to global warming. A meta-analysis of
June 2018
11-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
19 N addition observations, ranging from 15.4 to 300 kg N/ha/yr, found that N
enrichment increased wetland N2O emissions by 207% (Liu and Greaver. 2009).
11.3.1.2. Salt Marsh
New studies have evaluated the effects of N loading/N addition on tidal export, N
mineralization, nitrification, and microbial community structure in salt marshes. Tidal
export of N from salt marshes was studied in a long-term (>30 years) fertilization
experiment in Massachusetts (Brin et al.. 2010) in which tidal export of N increased with
increasing N load. Tidal export of ammonia increased linearly with increasing annual N
load, and tidal export of nitrate increased exponentially with increasing annual load (see
Table 11-2 for equations). Likewise, a linear relationship was documented between
increasing N and decreasing potential N mineralization after 7 months of N addition to
three salt marshes in California [see Table 11-2 for equation; (Vivanco et al.. 2015)1. In
this same study, there was no relationship between N addition and net nitrification rates;
however, the nitrification rates differed among the marsh sites.
In Kirkpatrick Marsh, MD, N addition decreased N retention in marshes dominated by
native plants and by the invasive lineage of Phragmites australis. N addition of 250 kg
N/ha/yr decreased ecosystem N retention by 55%, driven largely by a 57% decrease in N
retention belowground, in roots, rhizomes, and soil (Pastore et al.. 2016). N addition also
altered the relative allocation of N to pools in marsh plants. N addition increased the total
pool of N stored in aboveground biomass through stimulatory effects on biomass
production (see Appendix 11.4.1). but decreased the total pool of N stored in
belowground biomass by 6%. N addition decreased the pool of N in belowground
biomass but also shifted the distribution of root N higher in the soil profile than in control
plots (to 15-25 cm depth from 40-50 cm depth), making it more vulnerable to leaching
and to microbial transformation: N addition increased pore water NH4 concentrations
120% and N2O flux out of soils by 220% (Pastore et al.. 2016). The invasive species P.
australis established roots at depths below the native plant community (see
Appendix 11.3.2.1.2). which increased mineralization deep in the peat profile to increase
pore water [NH4+] 9-72% at 40-70 cm (Mozdzer et al.. 2016). Like many terrestrial
invasive species, P. australis may respond to N loading by creating a positive feedback
loop of increasing soil N availability and expanding invasion. Increasing reactive N may
change the diversity and composition of microbial communities responsible for
denitrification and nitrification. Change in the microbial communities could change the
rate of the chemical reactions that microbes perform in the wetland. Increasing N loads
alter the abundance of denitrifying bacteria.
June 2018
11-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
In the long-term marsh fertilization experiment in a salt marsh in Massachusetts, nirS
DNA sequences were amplified to assess denitrifying bacteria (Bowcn et al.. 2013). The
plots sampled were fertilized with sewage sludge, which contained other nutrients,
metals, and organic material in addition to N, and N addition rates were 221, 655, or
1,966 kg N/ha/yr. With the addition of N, denitrifying bacteria communities became
more dissimilar, as widely distributed bacteria declined, and abundance and richness of
unique denitrifiers increased (Bowcn et al.. 2013). Lage etal. (2010) sampled sediments
associated with Spartina patens in a New England salt marsh. This study used the a mo A
DNA sequence in sediments to sample the ammonia-oxidizing bacteria, which produce
nitrites from ammonia in the first step of nitrification. The majority of species affected by
N addition were within a clade of Nitrosospira-like sequences found in other salt
marshes. N addition decreased evenness of distribution and altered composition,
specifically the relative abundance of taxa within the marine and estuarine
Nitrosospira-like clade, suggesting that there are fine-scale genetic differences to the
bacteria's ability to use nitrogen.
An incubation experiment of soils collected from different zones of salt marsh in
Yancheng Nature Reserve added ammonium or nitrate in equal amounts of N to assess
what effects N chemical species had on microbial cycling of N. N addition increased net
N mineralization in all marsh zones, with oxidized N increasing net N mineralization
16-29%, and reduced N increasing net mineralization 58-69% (Zhang et al.. 2016c). N
forms had similar rates upon net nitrification rates, as NOs addition increased
nitrification rates 34-54%, and NH4 addition increased nitrification rates 65-94%. N
addition also increased the temperature sensitivity of N mineralization (Q10) in low
marshes (Zhang et al.. 2016c). suggesting a synergistic stimulation between increased N
loading and temperature upon N release from salt marshes to the marine environment.
11.3.1.3. Mangrove
In mangrove ecosystems, N addition suppressed N fixation and increased denitrification
(Whigham et al.. 2009). Laboratory incubations showed that the denitrification rate was
15 times higher in soils that had received 100 kg N/ha/yr than in control soils (N load on
N deposition in control soils not reported). However, field measurements of N2O
emissions were 5.6 times higher in fertilized than in control soils, indicating incomplete
denitrification in fertilized soils. Nitrogen fixation rates were suppressed by 88% in
fertilized plots (Whigham et al.. 2009).
June 2018
11-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
11.3.1.4. Freshwater Tidal Marsh
The Madisonville Nutrient Plots on the Tchefuncte River, LA, have received N addition
of 0, 50, 200, or 1,200 kg N/ha/yr for 11 years. There was no measurable effect of N
addition upon total N, microbial N, soluble N, denitrification rates, or C cycling in the top
10 cm of soil (Stcinmullcr et al.. 2016). although there were effects upon aboveground
productivity, plant stoichiometry, and plant community (see Appendix 11.4.2.
Appendix 11.5.3. and Appendix 11.8.1.2).
11.3.1.5. Riparian or Intermittent Wetlands
New studies on riparian wetlands include N addition relationships to denitrification rates
and bacterial community composition. In a riparian wetland in Durham, NC, wetland
species' potential denitrification activity increased linearly with increasing soil total
inorganic N [see Table 11-2 for equation; (McGill et al.. 2010)1. In freshwater riparian
wetlands along the James River, VA, soils were collected and incubated with added
nitrogen. Nitrogen addition increased the abundance of denitrifying bacteria—quantified
as copies of the nirS sequence—by 541% when labile organic matter was also added
(Morrissev et al.. 2013). When N addition occurred simultaneously with the addition of
recalcitrant organic matter, denitrifying bacterial abundance decreased 96% with
increased N. Community composition of denitrifying bacteria shifted in response to N
addition, as did the composition of bacteria capable of dissimilatory nitrate reduction to
ammonia [quantified as copies of nrfA sequence; (Morrissev et al.. 2013)1.
In riparian forests, increased N loads altered the symbiotic association between A In us
incana ssp. tenuifolia (grey alder) and Frankia (actinorhizal, N-fixing bacteria). N
addition of 100 kg N/ha/yr decreased the density of Frankia-hosting root nodules by 62%
compared with nodules from unamended control plots. Nodule respiration rates were
28% lower under N fertilization, indicating lower rates of Frankia metabolism, and N
fixation in nodules was 31% lower than in controls (Ruess et al.. 2013).
A recent study of an alpine meadow and intermittent wetland on the Tibetan Plateau
found that adding 30 kg N/ha/yr changed the wetland from a net sink to a net source of
N2O emissions to the atmosphere, suggesting N addition stimulated microbial
denitrification (Wang et al.. 2017c).
June 2018
11-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
11.3.1.6.
Bog and Fen
In peat bog ecosystems, N addition decreases ecosystem retention of N and increases N
exports in surface water from hydrologically connected bogs. In Mer Bleue Bog, Ontario,
N addition decreased the N retention efficiency of the ecosystem, as measured by the
recovery of an 15N tracer. When N was added at the rate of 16 kg N/ha/yr from
2000-2007, a pulse of added 15N was recovered at lower rates from fertilized plots than
from unamended control plots (N deposition was 8 kg N/ha/yr), indicating a 71%
reduction in average retention efficiency across biomass pools (Xing etal.. 2011). A
recent study (samples taken in 2014) in this bog found increases in nitrate and ammonium
concentrations in response to N addition in peat, with N addition increasing |NOs |
71-169% and [NH/] 90-269%, suggesting that the living Sphagnum moss is no longer
retaining N but leaching it into the peat below (Pinsonncault et al.. 2016).
A European study of N pools in bogs found that N deposition changes the retention of N
within vascular plants and peat soil. Under N treatment equivalent to deposition of 2 kg
N/ha/yr, 15N recovery was 66% for mosses, 3.8% for shrubs, and 1.8% for graminoids.
Deposition of 47 kg/ha/yr changed 15N recovery to 12% for shrubs as well as increasing
the total N stored in shrubs by 360% compared to mesocosms receiving 2 kg N/ha/yr. 15N
recovery also declined in peat under N deposition, with 15N recovery of 60% under 2 kg
N and 15N recovery of 36% under 47 kg N (Zaiac and Blodau. 2016). Increases in plant N
uptake along with decreasing efficiency of plant N retention (see Appendix 11.5.5) can
lead to increases in dissolved organic N exports from bogs. A study in a NPK-fertilized
and limed (addition of CaCCh) peatland catchment in the U.K. found high levels of N
export (14 kg N/ha watershed/yr) from the wetland to the Kinder River, with over half of
the exported N in the form of dissolved organic nitrogen (Edokpa et al.. 2015).
The Whim bog experiment in Scotland added N as either ammonium or nitrate to
simulate wet deposition, which with ambient N deposition of 8 kg N/ha/yr resulted in
treatment N loads of 16, 32, and 64 kg N/ha/yr. Increasing NOs loading increased
concentrations of dissolved organic nitrogen (DON) in pore water, while increasing NH4+
loading increased cation leaching from moss into pore water, and increased DIN in pore
water at the highest N addition level (Chiwa et al.. 2016). N addition impaired the ability
of the Sphagnum moss mat to absorb and retain N, beginning at 32 kg N/ha/yr for
oxidized N and 64 kg N/ha/yr for reduced N addition. This study illustrates the different
effects of reduced and oxidized N upon bog ecosystems (see also, Appendix 11.5.5).
However, a recent analysis by (Wieder et al.. 2016) suggested that North American and
European Sphagnum species have different tolerances for and responses to N loading, so
critical loads for North American bogs should be inferred with caution from European
bogs.
June 2018
11-12
DRAFT: Do Not Cite or Quote

-------
1	Recent studies have documented high rates of nitrogen fixation by microbial diazotrophs
2	living on and in dead Sphagnum cells (Vile 2014 and Larmola 2014 in (van Den Elzen et
3	al.. 2017). Microbial N fixation rates were observed at N deposition of 25 kg N/ha/yr, and
4	added another 6 kg N/ha/yr to moss and peat (van Den Elzen et al.. 2017).
11.3.1.7. Summary Table
Table 11-2 New studies on nitrogen addition effects on nitrogen cycling in
wetlands.
Type of
Additions or Load
Biological and



Ecosystem
(kg N/ha/yr)
Chemical Effects
Study Site
Study Species
Reference
Agricultural
0.02-90,480 kg
Removal: wetland N
Global
Reanalysis of
Jordan et al.
wetlands;
N/ha/yr as N load,
removal (Nremoval) is

data from
(2011)
intertidal
mean load for each
proportional to N load

109 global

marshes;
wetland class:
(Nload).

wetlands

freshwater
agricultural: 426,
log(Nremovai) — 0.943 x log



bogs and
intertidal: 211,
(Nioad) - 0.033. N



marshes;
freshwater bogs
removal efficiency is 26%



freshwater
and marshes: 890,
higher in nontidal than



swamps;
freshwater
tidal wetlands.



other
swamps: 69, other:




wetlands
280 kg N/ha/yr
Deposition = not
reported




Wetlands
N addition
N addition increases N2O
Global
Meta-analysis of
Liu and Greaver
(natural and
experiments, N
emissions 207% across

data collected
(2009)
agricultural)
addition of 15.4 to
300 kg N/ha/yr
Deposition = not
wetlands (n = 19).

from North
America, South
America,


reported


Europe, and




Asia

Salt marsh
Addition = 180 kg
Tidal export of N
Great
Spartina
Brin et al. (2010)

N/ha/yr (as
increases with increasing
Sippewissett
alterni flora,


Milorganite, NPK
N addition (Nadd as kg
Marsh, MA
Spartina patens,


10-6-4), 520 kg
N/ha/yr).

and Distichlis


N/ha/yr (as urea or
For NH4+ export

spicata


as Milorganite),
(NHxexport, in kg




1,560 kg N/ha/yr
N/season): NHxexport




(as Milorganite)
= 0.00083 x Nadd +0.432.
For NO3" export
(NOxexport, in kg
N/season): NOxexport =
0 122 x 6-1"10018Nadd



June 2018
11-13
DRAFT: Do Not Cite or Quote

-------
Table 11-2 (Continued): New studies on nitrogen addition effects on nitrogen
cycling in wetlands.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Salt marsh
Addition = 0,100,
200, 400, 800,
1,600,
3,200 kg N/ha/yr as
urea
N deposition =
3-5 kg N/ha/yr as
reported in
Tonnesen et al.
(2007)
Mineralization: potential
net N mineralization
(Nmin) decreases in a
linear response to added
N (Nadd, in g N/m2/yr):
Nmin — —0.0015 x Nadd
0.022.
Nitrification: no significant
relationship.
Morro Bay
National
Estuary,
Carpinteria Salt
Marsh
Reserve,
Tijuana River
Reserve
Estuary, CA
Salicornia
depressa
(Salicornia
virginica) stands
Vivanco et al.
(2015)
Salt marsh
Addition: 50 mg
N/kg soil, as
(NH4)2S04 or
NaNOs
Ambient
deposition = not
specified
N addition increased net
N mineralization in all
three marsh zones. NO3"
addition increased net N
mineralization 16-29%,
and NH4+ addition
increased net N
mineralization by
58-69%. The most
predictive soil parameter
for net N mineralization
was labile C:labile N
(r = -0.85)
N addition increased net
nitrification in all three
marsh zones. NO3"
addition increased net
nitrification 34-54%, and
NH4+addition increased
net nitrification by
65-94%. The most
predictive soil parameter
for net nitrification was
labile C:labile N
(r = -0.81).
N addition increased Q10
values (stimulatory effect
of temperature upon net
N mineralization) in P.
australis and S.
alterniflora marsh zones.
Aerobic soil
incubations
constructed
from core
samples from
Yancheng
National Nature
Reserve, China
Monospecific
stands across
marsh elevation
gradient:
Suaeda salsa
(native) in high
marsh,
Phragmites
australis
(native) in low
marsh, Spartina
alterniflora
(invasive, native
to North
America) in
previously
unvegetated
mud flats
Zhang et al.
(2016c)
June 2018
11-14
DRAFT: Do Not Cite or Quote

-------
Table 11-2 (Continued): New studies on nitrogen addition effects on nitrogen
cycling in wetlands.
Biological and
Chemical Effects
Study Site Study Species
Reference
Total marsh retention
(AG biomass, BG
biomass, soil) of 15N label
was 55% lower in N
treatment, because
belowground (BG
biomass + soil) N
retention was 57% lower
in N treatment. Across
elevated CO2 treatments,
belowground N retention
was 51% lower in N
treatment.
N addition increased N
mass in AG biomass for
C4 grasses. N addition
increased N mass in total
plant biomass for the first
6 yr, but not for the last
2 yr of measurement.
N addition decreased N
mass of fine roots. N
addition decreased total
N BG by 6% by shifting N
mass across the soil
profile: N addition
increased BG N mass at
depths of 15-25 cm, but
decreased BG N mass at
depths of 40-50 cm.
N addition increased N2O
flux from plots by 220%
and pore water NH4+
concentrations increased
120%.
Kirkpatrick
Marsh, Rhode
River,
Edgewater, MD
C3
Schoenoplectus
americanus; C4
Spartina patens
and Distichlis
spicata
Pastore et al.
(2016)
N addition increased
pore water
concentrations of NH4+
by 9 and 72% at depths
of 40 and 80 cm,
respectively.
P. australis increased
SOM decomposition in
recalcitrant peat 1.8-1.9
times over rates in
unvegetated peat.
Mesocosms of
invasive (from
seeds) or
native (from
rhzome
fragments)
plants in
Kirkpatrick
Marsh, Rhode
River,
Edgewater, MD
Invasive
Phragmites
australis; native
community of
Spartina patens,
Schoenoplectus
americanus
Mozdzer et al.
(2016)
Type of Additions or Load
Ecosystem (kg N/ha/yr)
Intertidal Addition = 250 kg
marsh	N/ha/yr (as NH4CI)
since 2006
Ambient
deposition = not
reported
Average N loads to
Chesapeake Bay
are 140 kg N/ha/yr
15N tracer applied
2006, measure AG
2005-2013,
measured BG 2014
Intertidal Addition = 250 kg
marsh	N/ha/yr (as NH4CI)
Ambient
deposition = not
reported
Salt marsh Addition = 1,630 kg
N/ha/yr as NH4NO3
Deposition = not
reported
Bacterial community: N Scarborough
addition decreases	Marsh, ME
evenness of
p-proteobacteria Laqe et al. (2010)
containing
amoA gene in
sediments
associated with
Spartina patens
ammonia-oxidizing
bacterial community and
changes community
composition (p = 0.017).
June 2018
11-15
DRAFT: Do Not Cite or Quote

-------
Table 11-2 (Continued): New studies on nitrogen addition effects on nitrogen
cycling in wetlands.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Salt marsh
Addition = sewage
sludge. Low
fert = 221 kg
N/ha/yr, high
fert = 655 kg
N/ha/yr, very high
fert = 1,966 kg
N/ha/yr
Deposition = not
reported
Bacterial community:
ubiquitous nirS
sequences declined with
increasing N addition
(higher abundance and
richness of unique
denitrifying species in
higher N treatments).
Great
Sippewissett
Marsh,
Falmouth, MA
Denitrifying
bacteria in
sediment
(community)
Bowen et al.
(2013)
Mangrove
Addition = 100 kg
N/ha/yr
Deposition = not
reported
Denitrification:
denitrification rate (lab
incubation) was 15x
faster in N addition
sediments.
Indian River
Lagoon, FL
(Impoundment
SLC-24)
Avicennia
germinans and
associated
sediments
Whiaham et al.
(2009)
Tidal marsh
Addition = 0, 50,
200, and 1,200 kg
N/ha/yr (as
slow-release
methylene urea,
which releases
NH4+), in
combination with 0
or 131 kg P/ha/yr,
to mimic
Mississippi River
diversion N and P
loading rates
Ambient
deposition = not
specified
11 yr of N addition had
no significant effect on
total P, total N, microbial
N, potentially
mineralizable N, or
potential denitrification
rates in the top 10 cm of
marsh soil.
Madisonville
nutrient plots
(microtidal
pulses of
10 cm),
Tchefuncte
River, Lake
Pontchartrain
Estuary, LA
Sagittaria
I and folia,
Polygonum
punctatum,
Eleocharis fallas
Steinmuller et al.
(2016)
Riparian
wetland
Addition = n/a
Deposition = not
reported
Denitrification: potential Durham, NC
1, 4, or 8 f
denitrification activity in
species from
spring increases with
Carex crinita,
total inorganic N in soil
Carex lurida,
(y = 6.96x+ 20.52).
Scirpus

cyperinus,

Juncus effusus,

Panicum

virgatum,

Chasmanthium

latifolium,

Eupatorium

fistulosum,

Veronia

noveboracensis,

Asclepias

incarnata, and

Lobelia

cardinalis.
McGill et al. (2010)
June 2018
11-16
DRAFT: Do Not Cite or Quote

-------
Table 11-2 (Continued): New studies on nitrogen addition effects on nitrogen
cycling in wetlands.
Type of
Additions or Load
Biological and



Ecosystem
(kg N/ha/yr)
Chemical Effects
Study Site
Study Species
Reference
Riparian
Addition = 100 kg
Frankia nodule densities
Bonanza
Alnus incana
Ruess et al.
floodplain
N/ha/yr
decreased 62%. Nodule
Forest LTER,
ssp. tenuifolia
(2013)
successional
Deposition = not
N fixation declined 31%
AK
and associated

forest
reported
and nodule respiration

Frankia strains


declined 28%.



Riparian
Addition = soil
Nitrate addition increased
James River,
Soil microbial
Morrissev et al.
wetland
incubations with
DNF (copies nirS)
Charles City
communities
(2013)

0.5, 2, 4 mg N/g
abundance 541 % when
County, VA
involved in


wet sediment as
labile OM was present

denitrification


KNOs
and decreased DNF

(DNF, gene


Deposition = not
abundance 96% when

nirS) or


reported
recalcitrant OM was

dissimilatory


present. Community
composition of DNF and
DNRA shifted in
response to nitrate
addition.

nitrate reduction
to NH4+ (DNRA,
gene nrfA)

Seasonal
Addition: 30 kg
N addition significantly
Mesocosms at
Carex
Wana et al.
wetland,
N/ha/yr
increased N2O
Luanhaizi
pamirensis,
(2017c)
alpine
Ambient
emissions, changing
wetlands,
Carex

meadow
deposition = 8.7 to
plots from net sinks to
Tibetan
atrofusca,


13.8 kg N/ha/yr
net sources of N2O to the
Plateau
Hippuris


atmosphere. N addition
reduced the global
warming potential (sum
of CO2, CH4, and N2O
fluxes) of the plots from a
net source to a net sink.

vulgaris,
Triglochin
palustris, and
Heleocharis
spp.

Ombrotrophic
Addition = 16 kg
Retention efficiency: N
Mer Bleue Bog,
Bog plant
Xina et al. (2011)
peat bog
N/ha/yr as NH4NO3
addition decreased the
Ontario,
community:


N deposition = 8 kg
retention efficiency of
Canada
dwarf shrub


N/ha/yr as
ecosystem N pools (15N
(measured
species and


quantified by
tracer) by 71%.
2007)
mosses:


Turunen et al.


Sphagnum


(2004)


magellanicum,
Sphagnum
capillifolium,
and Polytrichum
strictum

June 2018
11-17
DRAFT: Do Not Cite or Quote

-------
Table 11-2 (Continued): New studies on nitrogen addition effects on nitrogen
cycling in wetlands.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Ombrotrophic
peatland
Addition = 16 kg
N/ha/yr (5N, or 5x
background
deposition), 32 kg
N/ha/yr (1 ON), or
64 kg N/ha/yr
(20N), all as
NH4NO3
Also, 64 kg N/ha/yr
as NH4+ only
(20N-NH4), and
64 kg N/ha/yr as
NO3" only
(2ON-NO3)
Ambient
deposition = 3 to
5 kg N/ha/yr, -50%
as NO3"
Extractable N in peat
increases with N
addition: [NO3"]
increases 71% (5N),
125% (10N), and 169%
(20N); and [NH4+]
increases 90% (5N),
197% (10N), and 268%
(20N).
Mer Bleue bog,
Ottawa,
Canada
Nonliving peat
in top 10 cm of
bog
Pinsonneault et al.
(2016)
Ombrotrophic
peatland
Ambient
deposition = 2 kg
N/ha/yr at DS;
12 kg N/ha/yr at
WM; 47 kg N/ha/yr
at FS
Mesocosms were
established with
NH4NO3 solutions
that mimicked
relative N
deposition levels
15N tracer was
added in 48
applications over
6 mo at a rate of
23 kg N/ha/yr
At DS, 15N recovery was
60% in peat (depth
5-40 cm), 66% in living
Sphagnum, 3.8% in
shrubs, 1.8% in
graminoids, and 0.01%
for DIN (NO3" and NH4+).
At 47 kg N/ha/yr, 15N
recovery was 36% in
peat, 12% in shrubs, and
0.4% for DIN, but
otherwise similar to
recovery for pools at DS.
Data from LV and CF
mesocosms are not
considered because
mesocosm N addition
levels (CF: 300%
increase over DS N
solution, LV: 700%
increase over DS) did not
reflect ambient N
deposition differences
(CF: 750% increase over
DS deposition, LV: 300%
increase over DS).
Mesocosms
constructed
using peat bog
cores from
sites in
northern and
western
Europe—
Degero
Stormyr,
Sweden (DS);
Fenn's,
Whixall, and
Bettisfield
Mosses NNR,
UK (WM);
Frolichshaier
Sattelmoor,
Germany(FS)
(Peat cores
also collected
at Little
Vildmose,
Denmark [LV];
and Cors
Fochno, Wales,
UK [CF])
Sphagnum
capillifolium, S.
fallax, S.
magellanicum,
S. papillosum,
S. pulchrum, S.
rubellum,
Andromeda
polifolia,
Calluna
vulgaris, Erica
tetralix, Rubus
chamaemorus,
Vaccinium
oxycoccos,
Eriophorum
vaginatum,
Eriophorum
angustifolium
Zaiac and Blodau
(2016)
June 2018
11-18
DRAFT: Do Not Cite or Quote

-------
Table 11-2 (Continued): New studies on nitrogen addition effects on nitrogen
cycling in wetlands.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Peatland
catchment
9.6% of the
catchment area
fertilized with NPK
at 19.5 kg N/ha in
summer
35% of the
catchment area is
limed, 1,000 kg
CaC03/ha in
summer
28 kg N/ha/yr
(Helliwell et al..
2007)
Positive correlation
(r= 0.44) between
annual fluxes of DIN and
DON in stormflow.
Negative correlations
between fall DON and
DIN during baseflow
(r= -0.73) and stormflow
(r= - 0.92).
Annual total dissolved
nitrogen flux in river was
14 kg N/ha/yr, of which
54% is DON.
Kinder River
catchment,
south
Pennines, U.K.
Water samples
from the Kinder
River outlet
Edokpa et al.
(2015)
Ombrotrophic Addition = 8, 24,
peatland and 56 kg N/ha/yr
as either NH4+ or
NO3" since 2002
Ambient deposition
total N = 8 kg
N/ha/yr, 3 kg
N/ha/yr as wet
NOx, 3 kg N/ha/yr
as wet NHx, and
2 kg N/ha/yr as dry
NHx
In the oxidized N addition Whim bog,
plots, increasing uptake Edinburgh,
of N by Sphagnum	Scotland
increased pore water
DON and pore water
anion deficit.
In the reduced N plots,
increasing uptake of N by
Sphagnum increased
pore water cation
(K+ + Mg2+ + Ca2+ + H+).
Pore water DIN
increased only in the
highest addition of NhV
(56 kg N/ha/yr).
Sphagnum
moss
(Sphagnum
capillifolium, a
hummock-
forming
species)
Heathland
community:
Calluna
vulgaris,
Eriophorum
vaginatum,
Hypnum
jutlandicum,
Pleurozium
schreberi, and
Cladonia
portentosa
Chiwa et al. (2016)
DNF = denitrification; DNRA = dissimilatory nitrate reduction to ammonium; fert = fertilizer; ha = hectare; kg = kilogram;
KNO3 = potassium nitrate; LTER = Long-Term Ecological Research; N = nitrogen; NH4+ = ammonium; NH4NO3 = ammonium nitrate;
NPK = nitrogen, phosphorus, potassium; S = sulfur; yr = year.
11.3.2. Soil Carbon Cycling
1	Wetlands can be hotspots for emissions of carbon dioxide and methane. High water levels
2	in wetland soils prevent the rapid decomposition of dead roots and fallen stems by
3	aerobic bacteria or fungi. Over long time periods (decades, centuries, or millennia), these
4	conditions result in large stores of carbon belowground in wetland systems. Methanogens
5	are able to decompose some of these anaerobic, saturated carbon stores to produce
6	methane, and in severe droughts, a larger microbial community can produce large pulses
7	of carbon dioxide by decomposing freshly aerated older organic carbon. Endpoints of
June 2018	11-19	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
methane and carbon dioxide emissions have important implications for the wetland
function of long-term carbon storage and regulation of atmospheric composition.
A meta-analysis that included wetlands with other nonforest ecosystems indicated no
effect of N deposition on overall net ecosystem exchange of carbon (see
Appendix 11.3.2.2.1). In other words, any gain in carbon capture by photosynthesis was
offset by ecosystem respiration and C leaching. There were not enough studies to
evaluate wetlands as a separate category.
11.3.2.1. Belowground Decomposition, Respiration, and Biomass
Respiration includes the release of carbon dioxide that can be measured in the air. Soil
respiration may be the result of root or microbial activity. The process of decomposition
often releases carbon dioxide from organic matter. Decomposers feed on dead organic
matter, breaking it down into smaller compounds such as carbon dioxide or methane,
water, and nutrients. Nitrogen may change the rate of respiration and decomposition.
Low decomposition in wetlands results in the buildup of dead plant material in the soil. In
relatively closed systems, such as rain-fed bogs, a nutrient-poor, high-organic-acid
ecosystem is produced. In open systems where flowing surface water supplies additional
nutrients and sediments, the growth of new root systems on top of undecomposed roots
can trap sediment and increase the height of the soil surface. This accretion of new
wetland soil depends on both plant growth and maintenance of old carbon stored in soil
organic matter. Wetland existence depends on the sum of these processes: if the sum of
new plant growth does not exceed the decomposition of the soil organic matter, the
wetland may subside until water levels rise above the soil surface. In this case, the
drowned wetland area converts to an aquatic system. Wetland accretion is particularly
important in estuarine and marine systems, where wave action and tides regularly wash
sediment and dead plant material out of the wetland into aquatic systems, where the
material nourishes aquatic food webs. Belowground endpoints, such as bulk density, soil
organic matter, and root and rhizome production, are all indicators of wetland health and
continued existence (Table 11-3).
11.3.2.1.1. Across Wetlands
A new study of 90 wetlands around the Gulf Coast from Florida to Texas found that
higher N in the soil correlated with lower soil bulk density | see Table l l -3 for equation;
(Ncstlcrodc et al.. 2014)1. Wetlands included both tidal and nontidal marshes and
swamps. Reactive N or atmospheric N deposition loads were not quantified. This result
June 2018
11-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
suggests that higher N in wetlands correlates with a reduction in wetland soil stability,
making wetlands more susceptible to erosion.
Salt Marsh
A new study confirms N loading increases carbon dioxide production from salt marsh
soils. The new study uses an older model developed by Wigand et al. (2003). the
Nitrogen Loading Model (NLM) for wetlands and estuaries of Narragansett Bay, RI, to
determine annual N loads for different watersheds (sum of Nr from runoff, deposition,
and municipal waste, 10.3-6,727 kg N/ha/yr) within the bay. Sampling of cordgrass
marshes [Spartina alterniflora and Spartinapatens; (Wigand et al.. 2009)1 within these
watersheds showed that soil respiration of CO2 increased linearly and at similar rates with
increasing annual N load in stands of S. alterniflora and S. patens. Decreases in sediment
soil %C and %N and in belowground biomass were also observed as soil respiration
increased (Wigand et al.. 2009). Two additional experiments evaluated how N addition
affects soil respiration (Anisfeld and Hill. 2012) and general CO2 (see. Wigand et al..
2015). In both cases, N addition increased these rates; however, the amount of N was
over 1,000 kg/ha/yr, too high a level to evaluate the effects of N deposition.
In the Great Sippewissett Marsh fertilization experiment, fertilization with 520 kg N
decreased long-term carbon storage in substrates by 31% (Turner et al.. 2009). The
stability of the marsh substrate (measured as the physical resistance to force) decreased
60% in urea-only fertilized plots. In contrast, N addition to S. alterniflora marshes
located in Cocodrie, LA did not affect the physical resistance of the surface marsh
substrate, and no changes at any profile were detected with the addition of 1,200 kg
N/ha/yr (Turner. 2011). This study found negative effects upon marsh physical resistance
at high rates of N addition (2,300-18,600 kg N/ha/yr) not relevant to evaluating the
effects of N deposition.
The stability of marsh peat soils depends in part upon coarse roots and rhizomes, which
marsh plants produce for physical support and as perennial storage organs. A long-term
study at Goat Island, SC found stimulatory effects of N addition upon Spartina
alterniflora coarse roots and rhizomes, and organic matter in peat (Wigand et al.. 2015;
Davev et al.. 2011). but the experimentally added N load was 4,200 kg N/ha/yr, too high
to infer N deposition effects. Each year, marsh plants produce a new set of fine roots to
acquire nutrients from the sediment and pore water. In Kirkpatrick Marsh, vegetation
consisted of Schoenoplectus americanus, Spartina patens, and Distichlis spicata. N
addition (250 kg N/ha/yr) decreased fine root production by 42 and 84% compared to
control plots in the 3rd and 4th years of the experiment (Langlcv and Mcgonigal. 2012.
2010). When S. americanus and S. patens were planted in mesocosms for a sea level by
11-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
nitrogen fertilization factorial experiment, results were different. There was no
belowground response in mesocosms at current sea levels to nitrogen addition (Langlev
ct al.. 2013). However, in a mesocosm mimicking a 10 cm rise in sea level, N addition
increased belowground biomass by 130%. In a mesocosm mimicking a productive marsh
elevated 15 cm above current sea level, N addition increased fine roots in mescosms by
20-40% (Langlev et al.. 2013). The long term study at Goat Island found inhibitory
effects of N addition upon fine root mass (Davcv etal.. 2011). but the experimentally
added N load was too high to infer N deposition effects.
North American salt marshes are currently being invaded by the European lineage of
Phragmites australis, which is identified as a noxious weed or banned from sale or
transport by six states (USDA. 2015b). A Kirkpatrick Marsh study compared establishing
P. australis mesoscosms to mesocosms of the native plant community described by
Langlev et al. (2013). and found that 250 kg N/ha/yr increased belowground biomass of
the invasive species (Mozdzcr et al.. 2016). The invasive species had a greater rooting
depth than the native community, and N addition increased root biomass 23-69% at
depths of 10 to 40 cm, giving P. australis a competitive advantage over the native
species. P. australis roots primed microbial activity and increased decomposition of
buried peat by 1.8-1.9 times rates in uncolonized peat (Mozdzcr et al.. 2016). creating a
positive feedback loop between increasing N and invasive species expansion. A large
scale eutrophication experiment conducted at Plum Island Estuary, MA added nitrogen
and phosphorus to tidal inflows to raise tidal nitrate concentrations to 15 times the
background N load (Deegan et al.. 2012). Although the annual nitrogen load cannot be
calculated, the results of this experiment illustrate the mechanisms by which N addition
destabilizes salt marshes. Belowground biomass decreased 33% compared to control
marshes, and this in turn altered drainage, resulting in 4% higher water content in creek
banks in enriched marshes. Creek bank with decreased stabilizing root mats and
increased water content were less stable, with increasing numbers and lengths of creek
bed fractures over time (see Table 11-3 in Appendix 11.3.2.1.6 for equation). Creek
banks were so destabilized by N addition that the number of slumps, or creek edges
sliding into creek beds, was 113% higher in enriched than control marsh, which increased
the creek channel width:depth ratio by 28%. The unvegetated area of exposed mud was
200% higher in the enriched marsh as a consequence of the marsh destabilization
(Deegan et al.. 2012).
Freshwater Tidal Marsh
In marshes, coarse roots and rhizomes are produced by plants to serve as physical
supports and as storage organs for nutrients and carbohydrates. In a freshwater tidal
11-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
marsh dominated by Zizaniopsis miliacea in the Altamaha River Estuary, GA, addition of
500 kg N/ha/yr decreased the biomass of live rhizomes by 71%, and the mass of
macro-organic matter (living + dead roots) decreased 33% (Ketetal.. 2011). An
N-loading experiment added 50, 200, or 1,200 kg N/ha/yr to a freshwater tidal marsh on
the Tchefuncte River, LA (Graham and Mendelssohn. 2016). Researchers used two
different sampling methods to assess belowground biomass. Addition of 200 or 1,200 kg
N/ha/yr reduced biomass of established roots by 51%. Addition of 1,200 kg N/ha/yr also
increased growth of new roots into empty soil 106% over control (Graham and
Mendelssohn. 2016). This suggests that high and very high levels of N loading to tidal
freshwater marshes may destabilize existing marshes, while very high N loading may
accelerate establishment or expansion of new marshes. However, N addition in this
experiment had no measurable effect upon organic matter content, bulk density, total C,
total P, or N cycling in the top 10 cm of soil (see Appendix 11.3.1.4) (Steinmuller etal..
2016V
Intermittent Wetland
An N addition study was established on the Tibetan Plateau in an alpine meadow that is
also an intermittent wetland. N addition of 30 kg N/ha/yr increased belowground biomass
26%. N addition also stimulated CO2 uptake 25% (Wang et al.. 2017c). which offset the
concurrent stimulation of N2O emissions (see Appendix 11.3.1.5).
Bog and Fen
In the 2008 ISA, soil respiration in bogs had been studied in European countries under a
natural gradient of atmospheric N deposition from 2 to 20 kg N/ha/yr. Studies found
enhanced decomposition rates for material accumulated under higher atmospheric N
supplies resulted in higher carbon dioxide emissions. There are three new studies on an
Ontario bog and two new studies from European systems.
The Mer Bleue Bog in Ontario is the site of two concurrent fertilization experiments. In
one experiment, treatments were initiated in 2000 with nitrogen addition of 16 kg
N/ha/yr. The second experiment was initiated in 2005 in separate plots, where
nitrogen-only treatments of 32 kg N/ha/yr and 64 kg N/ha/yr were applied. Each
experiment had separate control plots. N addition at the rate of 16 kg N/ha/yr decreased
the concentrations of CO2 in peat substrate in the 8th year of N addition (2007),
decreasing CO2 by 19% at a 5-cm depth and 14% at a 25-cm depth (Wendel et al.. 201IV
N addition at the rate of 64 kg N/ha/yr did not significantly alter ecosystem respiration
when assessed 7 years after fertilization started (2011), but due to changes in
11-23
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
photosynthesis and declines in cover (see Appendix 11.5.5 and Appendix 11.8.1.5). net
ecosystem exchange in the growing season was 46% lower than in unamended bog plots
(Larmolaet al.. 2013).
In the Mer Bleue ombrotrophic peat bog, 16 kg N/ha/yr added from 2000 to 2007
increased peat biomass 17% in the top 10 cm of the substrate. In addition, median
temperature at a 5-cm depth in the peat was 19% lower in plots that received
16 kg N/ha/yr than in control plots (Wendel et al.. 2011). which the authors posited was
due to increased shading by increasingly productive shrub species. In contrast, soil
temperatures were elevated deeper in the peat substrate; at 20-cm depth, average daily
temperature was 0.6°C higher in plots with 16 kg N/ha/yr added than in control plots
(Wcndcl etal.. 2011). More recently, peat samples from Mer Bleue suggest that N
addition of 16 kg N/ha/yr, along with the resultant increase in shrub production, has
increased soluble phenolics 17% in peat, with peat C:K ratio increasing 45%
(Pinsonncault et al.. 2016). indicating a shift towards more recalcitrant C in the peat as
well as K limitation of the bog (see Appendix 11.5.5). In terms of the microbial
community responsible for decomposition and respiration in the peat, N addition in all
amounts increased P and K limitation, as indicated by increasing soluble C:P and C:K
ratios and increasing activity of P and K acquiring enzymes. However, N addition also
increased microbial breakdown of carbon, as indicated by 62-97% increases in the
activity of (3-D-glucosaminidase (Pinsonncault et al.. 2016). A recently initiated addition
experiment in the same bog added 64 kg N/ha/yr as either NH4+ or NOs and found that
reduced N resulted in 15% higher soluble phenolics and suppressed enzyme activities
compared to oxidized N (Pinsonneault et al.. 2016). This result suggests that oxidized N
is more likely to enhance microbial decomposition of peat than an equivalent amount of
reduced N.
A recent study of mesocosms collected from European bogs under a natural gradient of N
deposition from 2 to 24 kg N/ha/yr showed that increasing N deposition correlated with
increased respiration (and increased gross primary productivity, and increased net C
uptake; see Appendix 11.4.5) up to 15 kg N/ha/yr, but that all C fluxes decreased by 40%
between 15 and 24 kg N/ha/yr (Estop-Aragones et al.. 2016). Importantly, N stimulation
of respiration continued under drought conditions, although there was no relationship
during drought between N deposition and GPP or net carbon uptake. Inference from this
study is limited by sample size; there was only one mesocsom per bog (Estop-Aragones
et al.. 2016). In a study of decomposition rates conducted in Eastern Europe, N deposition
altered decomposition rates in bogs. When a standard cellulose substrate was placed in all
bogs, decomposition mass loss was higher at the high N deposition (20-25 kg N/ha/yr)
Jizera Mountains site, 170% higher when placed in areas dominated by Sphagnum
rubrum, and 510% higher mass loss in areas dominated by S. magellanicum, than in the
11-24
DRAFT: Do Not Cite or Quote

-------
1	low N deposition (12.5 kg N/ha/yr) Jeseniky Mountains site (Jirousck et al.. 2015). When
2	S. rubrum samples were collected from both sites and decomposed in a common site,
3	moss collected from the high N site decomposed more slowly, with 30% less mass lost
4	than from moss collected at the low N site, indicating that N deposition increased the
5	recalcitrance of moss to decomposition (Jirousck etal.. 2015). In contrast, at the Whim
6	Bog N addition experiment, there was no effect of added N as either ammonium or nitrate
7	(treatment N loads of 16, 32, and 64 kg N/ha/yr) upon decomposition of senesced
8	Sphagnum moss (Manninen et al.. 2016). although there were effects upon living moss
9	and water quality (see Appendix 11.5.5 and Appendix 11.3.1.5).
11.3.2.1.6. Summary Table
Table 11-3 Loading effects upon belowground carbon cycling.
Additions or
Type of	Load (kg	Biological and Chemical
Ecosystem	N/ha/yr)	Effects	Study Site Study Species Reference
Estuarine N dep = not	Soil total N (Nson, as %N) 90 Gulf Coast Marsh	Nestlerode et
marsh,	reported	increased as soil bulk	wetlands, Texas community, soil al. (2014)
estuarine 3 ^p = nof	density (SBD, as g/cc) to Florida	and pore water
shrub swamp, rpnnrteH	decreased,	chemistry
palustrine	In(Nsoii) = -1.9233 *
marsh,	SBD+ 0.4165.
palustrine
swamp
Salt marsh Watershed total	In S. alterniflora stands, soil Narragansett Bare sediments Wiqand et al.
N load (10.3,	respiration (y) increased Bay, Rl	in Spartina (2009)
11.5, 27.4, 37.4,	linearly with N loading (x)	patens and
2,729.9, 3,661,	(y = 0.0006x + 2.04). Soil	Spartina
S dep = not
reported
June 2018
11-25
DRAFT: Do Not Cite or Quote

-------
Table 11-3 (Continued): Loading effects upon belowground carbon cycling.
Type of
Ecosystem
Additions or
Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site Study Species Reference
Salt marsh
Addition:
1,050 kg N/ha/yr
(low) as NaNC>3,
2,100 kg N/ha/yr
(medium) as
NH4NO3 or
NaNOs, 4,200 kg
N/ha/yr (high) as
NH4NO3 in same
plots in over
several years
N dep = not
reported
S dep = not
reported
Medium N increased annual
sediment respiration
(g C/m2/yr) by 53%.
Hoadley Creek
Marsh, Guilford,
CT
Spartina
alterni flora
Anisfeld and
Hill (2012)
Salt marsh
520 kg N/ha/yr in
U (N as urea
treatment), UP
(urea + 208 kg
P/ha/yr as
phosphate), and
HF (milorganite,
10-6-4 NPK)
plots, as reported
by Brin et al.
(2010)
N dep = not
reported
S dep = not
reported
Fertilization decreased the
organic:inorganic ratio by
31% in older, deeper
(pre-1963) marsh substrate,
and decreased the physical
resistance of the marsh
substrate (shear vane
strength) by 60% in U plots,
36% in UP plots, and 35%
in HF plots.
Great
Sippewissett
Marsh, MA
Spartina
alterni flora,
Spartina
patens, and
Distichlis
spicata
Turner et al.
(2009)
Coastal salt
marsh
Addition:
4,200 kg N/ha/yr
as NH4NO3 or
(NH4)2S04
N dep = not
reported
S dep = not
reported
N addition increased coarse
root volume 63% in the top
10 cm, or 61% in the top
20 cm of sediment. Organic
matter increased 20% in the
top 21 cm of sediment. CO2
emission increased 41%.
Goat Island,
(measured
2008)
SC
Spartina
alterni flora and
associated
sediments
Wiaand et al.
(2015)
Salt marsh
4,200 kg N/ha/yr
as NH4NO3
N dep = not
reported
S dep = not
reported
N addition increased peat
mass fraction by 28% and
root/rhizome mass fraction
by 58% in wet shallow
(<10 cm) sediments. Fine
root mass decreased 39%
in shallow sediments.
Goat Island, SC
(measured
2008)
Spartina
alterni flora
Davev et al.
(2011)
June 2018
11-26
DRAFT: Do Not Cite or Quote

-------
Table 11-3 (Continued): Loading effects upon belowground carbon cycling.
Type of
Ecosystem
Additions or
Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site Study Species Reference
Estuarine salt
marsh
250 kg N/ha/yr
N dep = not
reported
S dep = not
reported
Total fine root production
decreased by 42 and 84%
in the 3rd and 4th yr.
Kirkpatrick
Marsh, MD
(measured in
3rd and 4th yr,
2008-2009)
Schoenoplectus Lanqlev and
americanus Meaoniaal
(C3), Spartina
patens (C4),
and Distichlis
spicata (C4)
(2012. 2010)
Estuarine salt
marsh
NH4CI solution
injected into
mesocosm peat
at root-level,
250 kgN/ha,
increases N by
40% above
annual average
background
concentration
Deposition not
reported
At increased marsh
elevation (15 cm above
current sea level), N
addition increased total fine
root mass by 20-40%.
At 10 cm sea level rise, N
addition increases BG
biomass by 130% under
ambient [CO2].
Kirkpatrick
Marsh, MD
Factorial
mesocosm
experiment with
varying sea
levels: 35 or
15 cm rise in
marsh
elevation;
current sea
level; 10, 20, or
30 cm rise in
sea level.
Mesocosms
planted with
2 rhizomes
Schoenoplectus
americanus and
10 stems
Spartina
patens;
harvested each
year
Lanqlev et al.
(2013)
Salt marsh
Addition = 250 kg
N/ha/yr (as
NH4CI)
Ambient
deposition = not
reported
N addition did not
significantly affect the
rooting depth of Phragmites
australis, which was
significantly deeper than
rooting depth of native
community.
N addition increased
belowground biomass of P.
australis 37-69% between
10 and 40 cm depth, with
peak increases at 10 cm
depth.
N addition increased pore
water concentrations of NH4
by 9 and 72% at depths of
40 and 80 cm, respectively.
P. australis increased SOM
decomposition in
recalcitrant peat 1.8-1.9x
over rates in unvegetated
peat.
Mesocosms of
invasive (from
seeds) or native
(from rhzome
fragments)
plants in
Kirkpatrick
Marsh, Rhode
River,
Edgewater, MD
Invasive
Phragmites
australis; native
community of
Spartina
patens,
Schoenoplectus
americanus
Mozdzer et al.
(2016)
June 2018
11-27
DRAFT: Do Not Cite or Quote

-------
Table 11-3 (Continued): Loading effects upon belowground carbon cycling.
Type of
Ecosystem
Additions or
Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site Study Species Reference
Salt marsh
Nitrate and
phosphate
dissolved in tidal
inflows to raise
aqueous NO3"
concentrations to
70-100 |jM and
raise PO43" to
5-7 |jM.
Background load
in tides: 5 |jM
NO3-, 1 |JM
PO43-.
N dep = not
reported
S dep = not
reported
Marsh stability (F, as creek
bank fractures) decreases
with cumulative N
enrichment (y, in years of N
enrichment):
F = 0.42y+ 0.79. Creek
fractures were 4.4* as long
in enriched as in control
marsh.
Water content in creek bank
increases 4%.
Number of bank slumps into
creek bed is 2.1 * higher in
enriched marsh. Channel
width:depth increased 28%.
Belowground biomass
decreases 33%.
Area of unvegetated
exposed mud increased
200%.
Plum Island
Estuary, MA
Primary tidal
creeks with
Spartina
alterniflora in
low marsh
along creeks,
and Spartina
patens in high
marsh platforms
Deeqan et al.
(2012)
Freshwater
marsh
500 kg N/ha/yr as Addition of 500 kg N/ha/yr Altamaha
NH4CI or urea
N dep = not
reported
S dep = not
reported
decreased the biomass of
live rhizomes by 71%, and
the mass of macro-organic
matter (living + dead roots)
decreased 33%.
Estuary, GA
(2007, 2008)
Zizaniopsis
miliacea,
Pontederia
cord at a, and
Sagittaria
land folia
Ket et al.
(2011)
Tidal marsh
Addition = 0, 50,
200, 1,200 kg
N/ha/yr as urea
Ambient
deposition = not
reported
Addition of 1,200 kg N/ha/yr
increased root colonization
of soil 106% over control
(0 kg added N) as
evaluated by living root
biomass within ingrowth
cores (p-value = 0.03).
Addition of 200 or 1,200 kg
N/ha/yr reduced biomass of
established roots by 51% as
evaluated by sampling
standing root biomass
(p-value = 0.02).
Tchefuncte
River, Lake
Pontchartrain,
LA
Sagittaria
land folia
Graham and
Mendelssohn
(2016)
June 2018
11-28
DRAFT: Do Not Cite or Quote

-------
Table 11-3 (Continued): Loading effects upon belowground carbon cycling.
Type of
Ecosystem
Additions or
Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site Study Species Reference
Tidal marsh
Addition = 0, 50,
200, and
1,200 kg N/ha/yr
(as slow-release
methylene urea,
which releases
NH4+), in
combination with
0 or 131 kg
P/ha/yr, to mimic
Mississippi River
diversion N and P
loading rates
Ambient
deposition = not
specified
11 yr of N addition had no
significant effect on organic
matter content, bulk density,
total C, total P in the top
10 cm of marsh soil.
Madisonville
nutrient plots
(microtidal
pulses of
10 cm),
Tchefuncte
River, Lake
Pontchartrain
Estuary, LA
Sagittaria
lancifolia,
Polygonum
punctatum,
Eleocharis
fall as
Steinmuller et
al. (2016)
Seasonal
wetland, alpine
meadow
Addition: 30 kg
N/ha/yr
Ambient
deposition = 8.7
to 13.8 kg N/ha/yr
N addition increased
belowground biomass by
26%.
N addition increased CO2
uptake by 25%. N addition
reduced the global warming
potential (sum of CO2, CH4,
and N2O fluxes) of the plots
from a net source to a net
sink.
Mesocosms at
Luanhaizi
wetlands,
Tibetan Plateau
Carex
pamirensis,
Carex
atrofusca,
Hippuris
vulgaris,
Triglochin
palustris, and
Heleocharis
spp.
Wang et al.
(2017c)
Ombrotrophic
peat bog
16 (low), 32
(medium), or 64
(high) kg N/ha/yr
as N (NH4NO3) or
NPK(NH4N03
and KHPO4)
S dep = not
reported
N dep = 8 kg
N/ha/yr
Medium and high NPK
increased ecosystem
respiration in 8th yr by 24
and 32% respectively, and
decreased surface
temperature by 11 and
13%, respectively. In high
NPK, the water table was
42% closer to peat surface.
Mer Bleue Bog,
Ontario, Canada
Bog plant
community:
dwarf shrub
species and
mosses
Sphagnum
magellanicum,
Sphagnum
capillifolium,
and Polytrichum
strictum
Juutinen et al.
(2010)
Ombrotrophic Addition: 16 kg
peat bog	N/ha/yr as
NH4NO3
N dep = 8 kg
N/ha/yr as
quantified by
Turunen et al.
(2004)
N addition lowered
substrate temperature at
5-cm depth, decreasing
median temperature by
19% and increased
substrate temperature at
20-cm depth, by an average
of 0.6°C.
N addition decreased CO2
concentrations in substrate
by 19% at 5-cm depth and
by 14% at 25-cm depth.
N addition increased peat
biomass 17%.
Mer Bleue Bog,
Ontario, Canada
(measured
2007)
Bog plant
community:
dwarf shrub
species and
mosses
Sphagnum
magellanicum,
Sphagnum
capillifolium,
and Polytrichum
strictum
Wendel et al.
(2011)
June 2018
11-29
DRAFT: Do Not Cite or Quote

-------
Table 11-3 (Continued): Loading effects upon belowground carbon cycling.

Additions or




Type of
Load (kg
Biological and Chemical



Ecosystem
N/ha/yr)
Effects
Study Site
Study Species
Reference
Ombrotrophic
16 (low), 32
Under high N, growing
Mer Bleue Bog,
Shrub species
Larmola et al.
peat bog
(medium), or 64
season net ecosystem
Ontario, Canada
(Vaccinium
(2013)

(high) kg N/ha/yr
exchange declined by 46%.

myrtilloides,


as N (NH4NO3) or


Ledum


NPK(NH4N03


groenlandicum,


and KH2PO4)


Chamaedaphne





calyculata) and





mosses





(Sphagnum





magellanicum,





Sphagnum





capillifolium,





Polytrichum





strictum)

June 2018
11-30
DRAFT: Do Not Cite or Quote

-------
Table 11-3 (Continued): Loading effects upon belowground carbon cycling.
Type of
Ecosystem
Additions or
Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site Study Species Reference
Ombrotrophic
peatland
Addition = 16 kg
N/ha/yr (5N, or
5x background
deposition), 32 kg
N/ha/yr (1 ON), or
64 kg N/ha/yr
(20N), all as
NH4NO3
Also, 64 kg
N/ha/yr as NhU"1"
only (20N-NH4),
and 64 kg N/ha/yr
as NO3" only
(2ON-NO3)
Ambient
deposition = 3 to
5 kg N/ha/yr,
-50% as NO3-
16 kg N/ha/yr (5N): N
addition in this amount
increased the peat C:K ratio
45% and increased
peat-soluble phenolics 17%.
In terms of microbial
activity, N addition
decreased soluble C:N
17%, and decreased NAG
activity by 56%. N addition
increased soluble C:P 33%
and phosphatase activity
7%. N addition increased
soluble C:K45%. N addition
increased BDG activity 62%
and phenol oxidase activity
12%.
32 kg N/ha/yr (10N) and
64 kg N/ha/yr (20N): N
addition in these amounts
altered nutrient ratios of
peat: 15-23% decrease in
C:N, 15% increase in C:P
(64 kg N only), and
72-109% increase in C:K
compared to control plots.
In terms of microbial
activity, N addition
decreased soluble C:N
30-42%,	and decreased
NAG activity 72-77%. N
addition increased soluble
C:P 36-39% and
phosphatase activity
41-49%. N addition
increased soluble C:K
32-51%. N addition
increased BDG activity
80-97% but decreased
phenol oxidase activity
31-63%.
In comparing the form of N
in addition (20N-NH4+ vs.
2ON-NO3"), peat soluble
phenolics were 15% higher
under NH4+ than under NO3"
. Activity of all enzymes was
lower under NHV compared
to NO3" addition: BDG
decreased 11%, NAG
decreased 32%,
phosphatase decreased
7%, and phenol oxidase
decreased 56%.
Mer Bleue bog,
Ottawa, Canada
Nonliving peat
in top 10 cm of
bog
Pinsonneault
etal. (2016)
June 2018
11-31
DRAFT: Do Not Cite or Quote

-------
Table 11-3 (Continued): Loading effects upon belowground carbon cycling.
Type of
Ecosystem
Additions or
Load (kg
N/ha/yr)
Biological and Chemical
Effects
Study Site Study Species Reference
Ombrotrophic
peatland
Ambient
deposition of
NOx, NHy, and
SOx:
1.95-24.07 kg
N/ha/yr and
3.18-36.90 kg
S/ha/yr
Deposition
estimates from
EMEP (European
Monitoring and
Evaluation
Programme)-
based IDEM
(Integrated
Deposition
Model) from
Pieterse et al.
(2007)
N and S deposition rates
are correlated across these
sites (r= 0.87).
Respiration (CresP, as g
C/m2/day) was positively
correlated with N deposition
(Ndep, as kg N/ha/yr) up to
15 kg N/ha/yr: CresP
= 0.448 + 0.067 x NdeP;
R2 = 0.61. The relationship
between N and respiration
was weaker during drought
(R2 = 0.38) and
post-drought (R2 = 0.50)
periods.
Respiration, GPP, and net
carbon uptake were all 40%
lower in the mesocosm from
the 24 kg N/ha/yr than from
the 15 kg N/ha/yr site.
Lab incubations
of 14 Sphagnum
mesocosms
collected in
saturated
hollows in
European bogs:
U.K. and Ireland
(n = 9), Poland
(n = 4) and
Slovakia (n = 1)
Mesocsoms
were maintained
at a high water
table (0-5 cm in
depth) for
40 days, then
were subject to
drought for
100 days, then
were re wetted to
high water table
(0-5 cm in
depth) for
200 days.
Sphagnum
fall ax
Estop-
Araaones et
al. (2016)
Ombrotrophic
bog
Wet + dry
deposition,
estimated by
Jirousek et al.
(2011)
Jireza: 20-25 kg
N/ha/yr (high N)
Jeseniky: 12.5 kg
N/ha/yr (low N)
Decomposition of a
standard substrate
(cellulose) is higher at high
N site, 170% higher mass
loss in S. rubellum zone
and 510% higher mass loss
in S. magellanicum zone.
In comparing decomposition
of S. rubellum grown at high
N or low N sites, S.
rubellum from high N mass
loss was 30% lower
(i.e., less decomposition).
Two sites:
Jizerka in Jireza
Mountains
(warm
suboceanic
climate) and
Vozka in
Jeseniky
Mountains (2°C
colder)
Sphagnum
fall ax,
Sphagnum
magellanicum,
and Sphagnum
rubellum/
russowii
Jirousek et al.
(2015)
Ombrotrophic
peatland
Addition = 8, 24,
and 56 kg N/ha/yr
as either NH4+ or
NOs"
Ambient
deposition = 8 kg
N/ha/yr
Decomposition rates of
Sphagnum were not
affected by elevated N
levels. Only the highest
level of N addition altered
pore water chemistry.
Whim bog,
Edinburgh,
Scotland
Sphagnum
moss
(Sphagnum
capillifolium)
Manninen et
al. (2016)
ANPP = aboveground net primary productivity; C = carbon; cm = centimeter; C02 = carbon dioxide; dep = deposition; g = gram;
ha = hectare; kg = kilogram; m = meter; N = nitrogen; NaN03 = sodium nitrate; NH4CI = ammonium chloride; NH4N03 = ammonium
nitrate; (NH4)2S04 = ammonium sulfate; NPK = nitrogen, phosphorus, potassium; S = sulfur; yr = year.
June 2018
11-32
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
11.3.2.2. Methane Emissions
Methane (CH4) is an important greenhouse gas that is over 20 times more effective at
trapping heat than carbon dioxide. The primary biological source of methane is microbial
(methanogens in the domain Archaea), as is the primary biological sink (methanotrophs
among the Bacteria and Archaea). N addition can stimulate methane flux from wetlands
by increasing methanogen activity when dissolved labile organic C is abundant (Kim et
al.. 2015b). or by decreasing methane oxidation by methanotrophs, as was recently
observed in a boreal peatland (Lozanovska et al.. 2016). Therefore, understanding the
controls on these microorganisms is important for predicting methane flux from
ecosystems. In terms of carbon emissions, the 2008 ISA reported that N deposition alters
CH4 flux in wetland and forested ecosystems. A meta-analysis of 25 N addition
observations (N addition 30 to 240 kg N/ha/yr) found that N addition increased CH4
emissions by 95% from wetlands and grasslands [see Table 11-4 for other nonsignificant
results; (Liu and Greaver. 2009)1.
There is new evidence that N loading increases methane production in soils. In the N
addition experiments replicated in three California marshes (Vivanco et al.. 2015; Irvine
et al.. 2012). field-measured methane flux from soils increased linearly with increasing N
load (0, 100, 200, 400, 800, 1,600, 3,200 kg N/ha/yr), such that CH4 flux from soils
increased by 1.23 |ig CH4/m2/day for each 10 kg N/ha/yr added R2 = 0.23 and P = 0.025
(Vivanco et al.. 2015; Irvine et al.. 2012). Methane flux increased from low or nearly
negative to net positive values above -100 kg N/ha/yr, showing that the effect of added N
on methanogenesis, the final step in anaerobic decomposition, offset any increase in
methanotrophy. Differential effects of N on both methanogenesis and methanotrophy
have been documented in other ecosystems (Irvine et al.. 2012; Liu and Greaver. 2009).
The linearity of the trend became less robust when the exposure time increased to
14 months.
Soils were collected from the Tijuana River Reserve near the sites used in Vivanco et al.
(2015) and incubated in a laboratory microcosm experiment with factorial C and N
additions (Irvine et al.. 2012). Adding C and N together increased methane production by
44% above controls (Irvine et al.. 2012). suggesting that N loading will increase
microbial methane production disproportionally when accompanied by pulses of labile C.
The authors concluded that temperate salt marshes generally emit low levels of methane,
but these values are also highly spatially and seasonally variable (Cheng et al.. 2010;
Magenheimer et al.. 1996; Bartlett et al.. 1985; King and Wiebe. 1978). Despite the high
variability observed, however, the field experiments suggest that increased N availability
June 2018
11-33
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
stimulated methanogenesis and increased methane emissions in southern California salt
marshes, as in other ecosystems (Liu and Greaver. 2009).
A study of a salt marsh in the Yellow River Delta, China, added N loads in different
chemical forms to the marsh to determine the effects of N form upon microbial
communities involved in C cycling. N was added as 50 kg N/ha/yr as NH44", 50 kg
N/ha/yr as NOs . or 100 kg N/ha/yr as NH4NO3. All three forms of N increased the
abundance of archaeal methanogens as well as the abundance of bacteria that form
syntrophies that stimulate methanogen metabolism (Xiao et al.. 2017). Ordination of
bacterial community composition suggested that all forms of N shifted the microbial
community in the same direction, but that reduced N had much stronger effects on
community composition than did oxidized N. N form also affected methane production
and fluxes: all N forms increased methane in pore water, but only NH44" and NH4NO3
additions increased methane emissions from the plots (Xiao et al.. 2017). Reduced forms
of N have stronger effects on microbial C cycling than do equivalent amounts of oxidized
Table 11-4 Nitrogen loading effects upon methane emissions.
N.
11.3.2.2.1. Summary Table
Type of Additions or Load
Ecosystem (kg N/ha/yr)
Biological and Chemical
Effects
Study
Study Site Species
Reference
Wetlands N addition
No effect across wetlands Global
(n = 6) of N addition on net
ecosystem exchange.
N addition increased CH4
emissions by 95% across
grass + wetland + anaerobic
agricultural systems (n = 25).
No effect across drained
wetlands of N addition (n = 6)
upon biological CH4 uptake.
Meta-	Liu and Greaver
analysis of (2009)
data
collected
from North
America,
South
America,
Europe, and
Asia
(natural experiments, N
and	addition of 10 to
agricultural) 562 kg N/ha/yr
Deposition = not
reported
Salt marsh 100,200,400,800,
Methane flux (y, as mg
CH4/m2/day) increased
linearly with N addition (A/add,
as g N/m2/yr):
y= 0.00123/Vadd- 0.0122.
Morro Bay, Marsh	Irvine et al. (2012)
Carpinteria dominated by
1,600, and 3,200 kg
N/ha/yr as granular
urea
Salt Marsh,	Salicornia
Tijuana	depressa
River	(formerly
Reserve,	Salicornia
CA	virginica)
June 2018
11-34
DRAFT: Do Not Cite or Quote

-------
Table 11-4 (Continued) Nitrogen loading effects upon methane emissions.
Type of
Additions or Load
Biological and Chemical

Study

Ecosystem
(kg N/ha/yr)
Effects
Study Site
Species
Reference
Salt marsh
0, 100, 200, 400,
CH4 flux from soils increased
Morro Bay
Stands of
Vivanco et al.

800, 1,600,
by 1.23 |jg CH4/m2/day for
National
Salicornia
(2015)

3,200 kg N/ha/yr as
each 10 kg N/ha/yr added
Estuary,
depressa


urea

Carpinteria
(formerly


N dep = 3-5 kg

Salt Marsh
Salicornia


N/ha/yr as reported

Reserve,
virginica)


in Tonnesen et al.

Tijuana



(2007)

River




Reserve
Estuary,
CA


Salt marsh
Addition of 50 kg
reduced N/ha/yr as
NH4CI, 50 kg
oxidized N as
KNO3, or 100 kg
N/ha/yr as NH4NO3
Ambient
deposition = not
specified
Addition of any form of N
increased bacterial
abundance of Geobacillus and
Clostridium, and increased
archaeal abundance of
Methanocellaceae.
Reduced-N and NH4NO3
additions decreased bacterial
abundance of Flavobacterium,
Bacillus, Gillisia,
Marinobacter, and
Desulfosarcina.
Redundancy analysis found
positive correlations between
ammonium-N measured in
sediment, methane flux, pore
water methane
concentrations, Geobacillus
abundance, Clostridium
abundance, and
Methanocellaceae
abundance.
In dry season, only 50 kg
reduced N addition plots are
methane sources; all other
plots are methane sinks.
In wet season, 50 kg oxidized
N increased pore water Chk
50 kg reduced N and 100 kg N
increased pore water CH4 as
well as CH4 emissions from
plots.
Marsh of
Phragmites
australis
and
Suaeda
heteroptera
in the
Yellow
River
Delta,
China
Bacterial and
archaeal
communities
assessed by
bacterial
primers
Ba338fand
Ba806r, and
archaeal
primers
Ar515f and
Ar907r
Xiao etal. (2017)
June 2018
11-35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
11.4.
Production and Aboveground Biomass
Aboveground biomass is a measure of how much carbon is fixed by wetland plants.
Changes in this endpoint can affect food webs within the wetland, including dependent
migratory species. Changes in aboveground biomass will also be important for food webs
in rivers, lakes, or estuaries connected by surface water flow, because plant litter from
wetlands is an important base for food webs downstream. In the 2008 ISA, evidence from
Canadian and European peatlands showed that N deposition had negative or mixed
effects on Sphagnum productivity, depending on history of deposition. In Canadian
ombrotrophic peatlands experiencing deposition of 2.7-8.1 kg N/ha/yr, peat
accumulation increased with N deposition, but accumulation rates had slowed by 2004,
indicating a degree of N saturation. Coastal wetlands responded to N enrichment with
increased primary production, shifting microbial and plant communities and altering pore
water chemistry, although many of the studies in coastal wetlands used N enrichment
levels more similar to those of wastewater than atmospheric deposition. The new studies
published from 2008-2015 include work on tidal marsh, mangroves, and ombrotrophic
bogs.
11.4.1. Salt Marsh
In the 2008 ISA, primary production of plant species in coastal wetlands typically
increased with N addition; however, most studies applied fertilizer treatments that were
several orders of magnitude larger than atmospheric deposition (Darby and Turner. 2008;
Tyler etal.. 2007; Wigand et al.. 2003; Mendelssohn. 1979). N fertilization experiments
in salt marsh ecosystems show biomass stimulation from 6 to 413% with application rates
ranging from 7 to 3,120 kg N/ha/yr (U.S. EPA. 1993).
A number of new studies have evaluated N effects on production and biomass in coastal
wetlands. In Elkhorn Slough, the largest coastal marsh in California, addition of 150 kg
N/ha/yr increased S. pacifica biomass in two of the six experimental sites (Goldman
Martone and Wasson. 2008); a 36% increase over unamended plots occurred in a site
with unrestricted tidal flow, and a 53% increase occurred in a marsh where
impoundments restricted regular tidal exchange. Darby and Turner (2008) reported on a
nitrogen addition experiment at Cocodrie, LA, in which five loads of nitrogen were
applied to S. alterniflora marshes. Aboveground biomass increased linearly with
increases in N load, although the regressions were not reported. Additions of 230 and
465 kg N/ha/yr did not significantly raise aboveground (AG) biomass above unamended
marsh biomass, but 930 kg N/ha/yr increased AG biomass 122%, 1,860 kg N/ha/yr
June 2018
11-36
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
increased AG biomass 124%, and 3,720 kg N/ha/yr increased AG biomass 141%. A
nitrogen addition experiment with seven different N loads (0, 100, 200, 400, 800, 1,600,
and 3,200 kg N/ha/yr) was replicated at three different California estuarine reserves
[Morro Bay, Carpinteria Salt Marsh, and Tijuana River Marsh Reserve; (Vivanco ct al..
2015)1. Aboveground biomass of existing S. depressa increased linearly in response to N
addition across all three marshes and N addition levels starting at 100 kg N/ha/yr (see
Table 11-5 for equation, R2 = 0.58). When harvested plots were resampled, the regrowth
of aboveground biomass increased in a saturating response to increasing N load, with no
significant increases in biomass in response to N above 1,600 kg N/ha/yr (Table 11-5 for
equation). Similar saturation of plant response occurred in N tissue (see Appendix 11.5).
In the eutrophication experiment at Plum Island, MA, tidal nitrate enriched to 15 times
background concentrations, along with P additions, increased aboveground shoot biomass
and specific mass of the vascular plant community by 16 and 9%, respectively (Deeganet
al.. 2012). When the tidal N enrichment was repeated in later years without P additions,
added loads of 620 kg N/ha/yr or 1,200 kg N/ha/yr increased low marsh Spartina
alterniflora shoot specific mass 3 and 12%, respectively (Johnson et al.. 2016a). In the
high marsh at this site, added N loads were smaller: at 140 kg N/ha/yr, S. alterniflora and
Distichlis spicata increased shoot-specific mass 14 and 8% (Johnson et al.. 2016a). There
were also changes in plant architecture in the same plots (see Appendix 11.6.1). but no
effects on plant community. The Kirkpatrick Marsh in Maryland is the site of a nitrogen
fertilization experiment in a tidal marsh community consisting of Schoenoplectus
americanus, Spartina patens, and Distichlis spicata. Over the 4 years of the experiment,
nitrogen addition of 250 kg N/ha/yr increased total aboveground biomass by 47-79%
(Langlcv and Mcgonigal. 2010; Langlev et al.. 2009). When the biomass response was
partitioned by carbon fixation mechanism (C3 or C4 plants), biomass of C3 plants
declined with N addition to 81% in the 3rd year and 55% in the 4th year of initial
fertilized S. americanus biomass. Biomass of C4 plants S. patens and D. spicata
increased 2-51 times the control over the 4 years of the experiment (see
Appendix 11.8.1). or increased 129% in the 4th year over initial fertilized C4 biomass. In
a related experiment at Kirkpatrick Marsh, mesocosms were constructed to mimic
marshes at different sea levels, planted with Schoenoplectus americanus and Spartina
patens, and subjected to the same nitrogen treatment (Langlev et al.. 2013). In this
experiment, sea level had the strongest impact on biomass. In mesocosms set at the
current sea level, N addition increased total AG biomass by 25-75%. The effects of N
addition were stronger at other sea levels, but responses at those marsh surface elevations
were species-specific. In a scenario of a 10-cm sea level rise, N addition increased total
AG biomass 85-570% above biomass of similarly inundated unfertilized mesocosms.
This strong response was driven almost entirely by the response of S. americanus, which
at a 10-cm sea level rise increased its biomass 150-533% in response to N addition. In
11-37
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
mesocosms that mimicked a 15- or 35-cm increase in marsh elevation above sea level (as
can occur in a highly productive marsh where productivity + accretion >
decomposition + export), total AG biomass increased only 15-95%. This community
response was driven by the response of S. patens, which at a 15 - to 35-cm marsh
elevation increase responded to nitrogen addition with 55-145% increases in AG
biomass (Langlev et al.. 2013). These responses reflect the tolerance of these plants for
inundation by tides and suggest that N addition may increase the productivity only of
plants growing squarely within the range of their physiological tolerance.
Several other studies have evaluated N addition levels over 1,000 kg N/ha/yr; in all cases,
N addition increased biomass, but the high levels of experimental N addition
exponentially exceed N loading from deposition, limiting the scope of inference from this
literature (Anisfeld and Hill. 2012; Nelson and Zavaleta. 2012; Ryan and Bover. 2012).
11.4.2. Mangrove
Mangrove ecosystems are ecologically and economically important because they provide
habitat not just to a diversity of resident species but also serve as nurseries to the juvenile
lifestages of many marine fish species. As slow-growing tree species, mangroves are
rarely destructively sampled in order to estimate productivity and aboveground biomass.
The mangroves at Indian River Lagoon, FL are the site of several nitrogen addition
experiments. Whigham et al. (2009) added 100 kg N/ha/yr to plots of dwarf Avicennia
germinans, black mangrove, which increased productivity by increasing the number of
new branches 150% above control plots.
In a long-term (since 1997) N addition experiment that added approximately
11,200 kg N/yr to each tree, growth rate of Rhizophora mangle, quantified by measuring
shoot elongation rates increased 290% in the shoreline fringe zone and increased 1,340%
in the interior scrub zone (Feller et al.. 2009). At Merritt Island National Wildlife Refuge,
N addition of 1,400 kg N/ha/yr in the ecotone where mangroves and cordgrasses grow in
competition increased A germinans seedlings' production of new leaves by 42%, which
increased total leaf biomass by 72% over unamended trees (Simpson et al.. 2013) and
altered above- and below-ground stoichiometry (see Appendix 11.5.2).
11.4.3. Freshwater Tidal Marsh
In a freshwater tidal marsh on the Tchefuncte River, LA, Graham and Mendelssohn
(2010) conducted an N loading experiment that added 50, 200, or 1,200 kg N/ha/yr to a
plant community dominated by Sagittaria lancifolia, Eleocharis fallax, and Persicaria
June 2018
11-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
punctata (formerly Polygonum punctatum). The aboveground net primary productivity of
the community increased with increasing N in a negative quadratic function, with ANPP
not increasing above 200 kg N/ha/yr (Graham and Mendelssohn. 2010).
At the Altamaha River Estuary, GA, N addition experiments were conducted in tidal
freshwater marshes dominated by giant cutgrass Zizaniopsis miliacea. Addition of 500 kg
N/ha/yr increased aboveground biomass 1.4-3.8 times control biomass in the 2nd
through 5th years of fertilization (Ket et al.. 2011; Frost et al.. 2009). A nitrogen
experiment conducted in estuarine marshes and swamps on the Nanticoke River in the
Chesapeake Bay found that while an added load of 670 kg N/ha/yr did not significantly
change total herbaceous plant community aboveground biomass, it did increase biomass
of the invasive Typha spp. by 47% (Baldwin. 2013). There are several other new studies
on the effects ofN loading on aboveground biomass of freshwater tidal ecosystems;
however, the addition rates are greater than 500 kg N/ha/yr.
11.4.4. Intermittent Wetland
An N addition study was established in an alpine meadow, which is also an intermittent
wetland on the Tibetan Plateau. N addition of 30 kg N/ha/yr increased ANPP of the
graminoid and herbaceous plant community by 11.5% (Wang et al.. 2017c). which also
affected C and N fluxes of the system (see Appendix 11.3).
11.4.5. Bog and Fen
In Sphagnum-dominated ombrotrophic bogs, higher N deposition resulted in higher tissue
N concentrations and greater NPP (Aldous. 2002a). but lower bulk density. A study of
23	ombrotrophic peatlands in Canada with deposition levels ranging from 2.7 to 8.1 kg
N/ha/yr showed peat accumulation increasing linearly with N deposition; however, in
later years of the study, the rate had begun to slow, indicating limited capacity for N to
stimulate accumulation (Turunen et al.. 2004). A study of European bogs under a
deposition gradient of 2 to 24 kg N/ha/yr found that gross primary productivity and net
carbon uptake of Sphagnum fallax mesoscosms were positively correlated with increasing
N deposition up to 15 kg N/ha/yr, but that the mesocosm collected from a bog receiving
24	kg N/ha/yr had 40% lower GPP and net C uptake than the mesocosm receiving 15 kg
N/ha/yr (Estop-Aragones et al.. 2016). However, there was N stimulation of GPP and net
C uptake at deposition levels below 15 kg N/ha/yr when mesocsoms were subject to
drought conditions. Inference from this study is limited by sample size; there was only
one mesocsom per bog (Estop-Aragones et al.. 2016).
June 2018
11-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
In the Mer Bleue ombrotrophic peat bog, 8 years of N addition at the rate of
16 kg N/ha/yr increased biomass of mosses (Sphagnum magellanicum, Sphagnum
capillifolium, and Polytrichum strictum) by 30% of moss biomass in control plots (Xing
ctal.. 2011).
In oligotrophic bogs in Gogebic County, MI, N addition of 60 kg N/ha/yr increased total
plant community productivity by 82% over unamended plot productivity, resulting in
2.6 times as much biomass in fertilized plots as biomass in unamended plots (Iversenet
al.. 2010). The plant community trends reflected the responses of the dominant vascular
shrub Chamaedaphne calyculata, which increased its productivity by 87%, thus
increasing its biomass in N addition plots 105% over its biomass in control plots (Iversen
et al.. 2010V
In general, vascular plants are able to respond more rapidly to N deposition than are moss
species. A stable isotope N addition study in Bourtanger Moor, Germany, showed that
grass Lolium multiflorum and sedge Eriophorum vaginatum responded to increasing N
addition of approximately 30-55 kg N/ha/yr with linear increases in aboveground
biomass (Hurkuck et al.. 2015). Plant interception of N deposition also varied between
species and with size for L. multiflorum (see Appendix 5).
Relative changes in biomass can lead to plant community changes. Model results from
Logofet and Alexandrov (1984) suggest 7 kg N/ha/yr is the threshold for an oligotrophic
bog to become a mesotrophic fen dominated by trees, as found in the 1993 Oxides of
Nitrogen AQCD. In freshwater-rich fens in Gogebic County, MI, N addition of
60 kg N/ha/yr increased productivity of dominant grass Calamagrostis canadensis, which
increased the species' biomass to 600% over its biomass in unamended fen plots (Iversen
et al.. 2010).
In many bogs and fens, N addition increases biomass and then abundance of vascular
plants and N-tolerant moss species, while N-sensitive species decline. A study of purple
pitcher plant, Sarracenia purpurea, growing in bogs across a deposition gradient of
3.4-5.0 kg N/ha/yr in the Adirondack Mountains, NY, identified a threshold for plant
response to N (Crumley et al.. 2016). with plant growth depressed 8-21% at deposition
levels above 4.1 kg N/ha/yr, and with other negative effects on plant physiology (see
Appendix 11.5.5). In a European mesocosm study of four moss species, elevated N
concentrations typical of rainwater in the fen depress growth rates 18% compared to
mosses growing in N concentrations typical of the fen groundwater (Andersen et al..
2016). Additional experimental N loading further decreases growth rate of all four
species. Two rare moss species, with declining abundance over the last 100 years, show
high sensitivity to N, with more growth rate variation explained by N addition (18-23%)
11-40
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
than for the growth rate in the more common moss species (5-7%) (Andersen ct al..
2016V
Recent research has addressed differential responses of bogs and fens to reduced or
oxidized forms of N addition. In an Irish fen dominated by peat moss Sphagnum
contortum and brown moss Scorpidium revolvens, 50 kg N/ha/yr was applied as NOs, in
one treatment and as NH4+ in another treatment (Paulisscn et al.. 2016). Reduced N
decreased S. revolvens biomass 67% compared to control plots, while oxidized N had no
measurable effect on S. revolvens biomass. S.contortum biomass was not affected by
reduced or oxidized N additions, although the plant's physiology did change in response
to both treatments (see Appendix 11.5.5).
11.4.6. Summary Table
Table 11-5 Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects
Study Site Study Species Reference
Coastal high
marsh
Addition: 150 kg
N/ha/yr as urea
S dep = not
reported
N dep = not
reported
At two sites with
unrestricted and restricted
tidal flow, fertilization
increased S. pacifica
biomass by 36 and 53%,
respectively.
Six sites at
Elkhorn Slough,
Watsonville, CA
Marsh community
dominated by
Sarcocornia
pacifica, also
contained
Jaumea carnosa,
Frankenia salina,
Spergularia
salina, Distichlis
spicata, and
Atriplex
caiifornica/
triangularis
Goldman
Martone and
Wasson
(2008)
Estuarine salt
marsh
Addition: 100, 200,
400, 800, 1,600,
3,200 kg N/ha/yr as
urea
N dep = 3-5 kg
N/ha/yr as reported
in Tonnesen et al.
(2007)
Aboveground biomass
(AGB) increases in a
linear response to N
addition (Nadd, as g
N/m2/yr), AGB = 0.001 *
Nadd + 1.117 (R2 = 0.58),
while biomass regrowth
(AGR) increases in a
saturating response to N,
AGR =
-1.16 x e("001Nadd' + 1.89.
Morro Bay
National
Estuary,
Carpinteria Salt
Marsh Reserve,
Tijuana River
Reserve
Estuary, CA
Salicornia
depressa
(Salicornia
virginica) stands
Vivanco et al.
(2015)
June 2018
11-41
DRAFT: Do Not Cite or Quote

-------
Table 11-5 (Continued): Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects
Study Site Study Species Reference
Salt marsh
Nitrate and
phosphate
dissolved in tidal
inflows to raise
aqueous NO3"
concentrations to
70-100 |jM and
raise PO43" to
5-7 |jM.
Background load in
tides: 5 |jM NO3",
1 |jM PCM3".
N dep = not
reported
S dep = not
reported
Vascular plant shoot
biomass increased 16%,
shoot specific mass
increased 9%.
Plum Island
Estuary, MA
Primary tidal
creeks with
Spartina
alterniflora in low
marsh along
creeks, and
Spartina patens in
high marsh
platforms
Deeaan et al.
(2012)
Salt marsh
Addition of nitrate
dissolved in
incoming tide
2011-2012,
addition of N at
70-100 |jM NaNOs"
in tide for added
load of
620-1,200 kg
N/ha/yr in low
marsh and
70-140 kg N/ha/yr
in high marsh
Ambient
deposition = not
specified
Low marsh S. alterniflora
in enriched creeks
received an additional N
load of 1,200 kg N/ha/yr
in 2011. That year, N
addition increased
shoot-specific mass 12%
above S. alterniflora in
reference creek plots.
Low marsh S. alterniflora
in enriched creeks
received an additional N
load of 620 kg N/ha/yr in
2012. That year, N
addition increased
shoot-specific mass 3%.
High marsh S. alterniflora
and D. spicata in enriched
creeks received an
additional N load of
140 kg N/ha/yr in 2011. N
addition increased S.
alterniflora shoot-specific
mass 14%. In D. spicata,
N addition increased
shoot-specific mass 8%.
Plum Island
Sound Estuary,
MA
Spartina
alterniflora,
Distichlis spicata,
and Spartina
patens (high
marsh); Spartina
alterniflora (low
marsh)
Johnson et al.
(2016a)
Estuarine
marsh
Addition: 1,337 kg
N/ha/yr as urea.
S dep = not
reported
N dep = not
reported
In S. pacifica, biomass
increased 54-185%
across habitats.
China Camp
State Park, CA
Sarcocornia
pacifica dominant
(C3 succulent
shrub), Distichlis
spicata (C4
grass), and
Jaumea carnosa
(C3
semisucculent
forb)
Ryan and
Bover (2012)
June 2018
11-42
DRAFT: Do Not Cite or Quote

-------
Table 11-5 (Continued): Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects
Study Site Study Species Reference
Coastal salt Addition: 3,000 kg AG biomass increased 28 Elkhorn Slough, Sarcocornia
Nelson and
marsh	N/ha/yr as NH4NO3
S dep = not
reported
N dep = not
reported
and 216% in successive
summers, and increased
shoot-to-root ratio 249%.
Monterey Bay,
CA
pacifica dominant, Zavaleta
Distichlis spicata, (2012)
Frankenia salina,
and Jaumea
carnosa
Coastal salt
marsh
230, 465, 930,
1,860, or 3,720 kg
N/ha/yr as
(NH4)2S04
S dep = not
reported
N dep = not
reported
AG biomass increased
linearly with N load
(regression not given).
Aboveground live
biomass increased 122,
124, and 141% in
response to 930, 1,860,
and 3,720 kg N.
LUMCON,
Cocodrie, LA
Spartina
alterniflora
Darby and
Turner (2008)
Coastal salt
marsh
1,050 kg N/ha/yr
(low) as NaNC>3,
2,100 kg N/ha/yr
(medium) as
NH4NO3 or NaNOs,
4,200 kg N/ha/yr
(high) as NH4NO3 in
same plots in over
several years
S dep = not
reported
N dep = not
reported
ANPP increased by 132%
in low, 130% in medium,
and 120% in high N
treatments.
Hoadley Creek
Marsh, Guilford,
CT
Spartina
alterniflora
Anisfeld and
Hill (2012)
Estuarine salt Addition: 250 kg
marsh	N/ha/yr
S dep = not
reported
N dep = not
reported
S. americanus
aboveground biomass
decreased by 19 and
45% in the 3rd and 4th yr,
respectively.
Biomass of S. patens and
D. spicata increased
129% in the 4th yr over
initial fertilized C4
biomass.
Kirkpatrick
Marsh, MD
(measured in
3rd and 4th yr,
2008-2009)
Schoenoplectus
americanus (C3),
Spartina patens
(C4), and
Distichlis spicata
(C4)
Lanqlev and
Meqoniqal
(2012. 2010)
Estuarine salt Addition: 250 kg
marsh	N/ha/yr
S dep = not
reported
N dep = not
reported
In the first two growing
seasons, fertilization
increased aboveground
biomass in the second
season by 57%.
Kirkpatrick
Marsh, MD
(measured in
1st and
2nd years,
2006-2007)
Schoenoplectus
americanus,
Spartina patens,
and Distichlis
spicata
Lanqlev et al.
(2009)
June 2018
11-43
DRAFT: Do Not Cite or Quote

-------
Table 11-5 (Continued): Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects
Study Site Study Species Reference
Estuarine
marsh
NH4CI solution
injected into
mesocosm peat at
root level,
250 kgN/ha,
increases N by 40%
above annual
average
background
concentration
Deposition not
reported
At current sea level, N
addition increased total
AG biomass by 25-75%
in 1st yr.
With sea level rise 10 cm,
S. americanus AG
biomass increases
150-230% in 1 st yr,
110-533% in 2nd yr.
Total AG biomass at
10 cm increased 85-95%
in 1st yr, 100-570% in
2nd yr.
At increased marsh
elevation (15 and 35 cm
above sea level 1st yr,
15 cm 2nd yr), S. patens
AG biomass increases
55-145% in 1 st yr,
60-90% in 2nd yr. Total
AG biomass at these
elevations increased
15-95% in 1st yr,
30-35% in 2nd yr.
Kirkpatrick
Marsh, MD
Factorial
mesocosm
experiment with
varying sea
levels: 35- or
15-cm rise in
marsh elevation;
current sea level;
10-, 20-, or 30-cm
rise in sea level.
Mesocosms
planted with
2 rhizomes
Schoenoplectus
americanus and
10 stems Spartina
patens', harvested
each year.
Lanalev et al.
(2013)
Mangrove/
marsh
ecotone
Addition: 1,400 kg
N/ha/yr
S dep = not
reported
N dep = not
reported
Mangrove leaf production
increased by 42% and
leaf biomass increased by
72%.
Merritt Island Avicennia
National	germinans
Wildlife Refuge,
FL
(Impoundment
T9)
Simpson et al.
(2013)
Mangrove
Addition: 1,400 kg
N/ha/yr
S dep = not
reported
N dep = not
reported
Fertilization increases
growth rate (shoot
elongation) by 290% in
the fringe zone and by
1,340% in the scrub zone.
Indian River
Lagoon, FL
(impoundment
MI23)
Rhizophora
mangle
Feller et al.
(2009)
Mangrove
100 kg N/ha/yr
N addition increased the
number of new branches
150% in Avicennia.
Indian River
Lagoon, FL
(impoundment
SLC-24)
Avicennia
germinans and
associated
sediments
Whiqham et
al. (2009)
Estuarine
Addition: 670 kg
Typha spp. biomass
Nanticoke
Plant community
Baldwin
tidal marsh
N/ha/yr
increased by 95%.
River, MD and

(2013)

S dep = not

DE



reported





N dep = not





reported




June 2018
11-44
DRAFT: Do Not Cite or Quote

-------
Table 11-5 (Continued): Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects	Study Site Study Species Reference
Freshwater
50, 200, or 1,200 kg
Community ANPP
Tchefuncte
Oligohaline plant
Graham and
estuarine
N/ha/yr as
increased with medium N
River,
community
Mendelssohn
marsh
Nutralene
load (Nadd as kg N/ha/yr),
Madisonville,
dominated by
(2010)

methylene urea
ANPP = -0.00165 x Nadd2
LA
Sagittaria


S dep = not
reported
+ 2.5091 x Nadd + 1270.3.

lancifolia,




Eleocharis fallax,


N dep = not
reported


and Polygonum
punctatum

Freshwater
500 kg N/ha/yr as
N addition increased
Altamaha River,
Zizaniopsis
Frost et al.
tidal marsh
NH4CI or urea
aboveground biomass by
140% and plant height by
32% in Z. miliacea. N
addition decreased leaf
GA
miliacea
(2009)
[N] by 99% and leaf [P] by
25%.
Freshwater
marsh
500 kg N/ha/yr as
NH4CI or urea
N dep = not
reported
S dep = not
reported
In Z. miliacea,	Altamaha	Zizaniopsis
aboveground biomass	Estuary, GA miliacea,
increased by 2.9-3.8x the Pontederia
control, as leaf number cordata, and
increased 52% and plant Sagittaria
height increased 25-40%. lancifolia
Ket et al.
(2011)
Seasonal Addition: 30 kg N addition increased	Mesocosms at Carex pamirensis, Wang et al.
wetland, N/ha/yr	ANPP 11.5% across	Luanhaizi	Carex atrofusca, (2017c)
alpine	Ambient	average (3 cm above soil wetlands,	Hippuris vulgaris,
meadow deposition = 8 7 to surface) and lowered	Tibetan Plateau Triglochin
13 8 kg N/ha/yr (^0 cm below soil	palustris, and
surface) water table	Heleocharis spp.
levels.
June 2018	11-45	DRAFT: Do Not Cite or Quote

-------
Table 11-5 (Continued): Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects
Study Site Study Species Reference
Ombrotrophic
peatland
Ambient deposition
of NOx, NHy, and
SOx: 1.95-24.07 kg
N/ha/yr and
3.18-36.90 kg
S/ha/yr
Deposition
estimates from
EMEP (European
Monitoring and
Evaluation
Programme)-based
IDEM (Integrated
Deposition Model)
from Pieterse et al.
(2007)
N and S deposition rates
are correlated across
these sites (r = 0.87).
Gross primary
productivity (GPP, as g
C/m2/day) increased with
N deposition (Ndep, as kg
N/ha/yr) up to 15 kg
N/ha/yr: GPP= 1.262
+ 0.219NdeP; R2 = 0.65.
There was no relationship
between N deposition and
GPP during drought and
post-drought periods.
Net carbon uptake by
Sphagnum (Cuptake, as g
C/m2/day) was positively
correlated with N
deposition (Ndep, as kg
N/ha/yr) up to levels of
15 kg/ha/yr:
Cuptake—0.76 + 0.155xNdep,
R2 = 0.52. There was no
relationship between N
deposition and GPP
during drought and
post-drought periods.
Respiration, GPP, and
net carbon uptake were
all 40% lower in the
mesocosm from the 24 kg
N/ha/yr than from the
15 kg N/ha/yr site.
Lab incubations
of
14 Sphagnum
mesocosm
collected in
saturated
hollows in
European bogs:
U.K. and
Ireland (n = 9),
Poland (n = 4),
and Slovakia
(n = 1).
Mesocsoms
were
maintained at a
high water table
(0-5 cm depth)
for 40 days,
then were
subject to
drought for
100 days, then
were re wetted
to high water
table (0-5 cm
depth) for
200 days.
Sphagnum fallax
Estop-
Araqones et
al. (2016)
Ombrotrophic 16 (low), 32
peatbog (medium), or64
(high) kg N/ha/yr as
N (NH4NO3) or NPK
(NH4NO3 and
KHPO4)
In high NPK, shrub
biomass increased, with
40% higher leaf biomass,
and 86% higher woody
biomass.
Mer Bleue Bog, Bog plant
Juutinen et al.
Ontario,	community: dwarf (2010)
Canada	shrub species
and mosses
Sphagnum
magellanicum,
Sphagnum
capillifolium, and
Polytrichum
strictum
Ombrotrophic
peat bog
Addition: 16 kg
N/ha/yr as NH4NO3
N dep = 8 kg
N/ha/yr as
quantified by
Turunen et al.
(2004)
N addition decreased
aboveground moss
biomass by 30%.
Mer Bleue Bog, Bog plant
Xing et al.
Ontario,	community: dwarf (2011)
Canada	shrub species
(measured and mosses
2007)	Sphagnum
magellanicum,
Sphagnum
capillifolium, and
Polytrichum
strictum
June 2018
11-46
DRAFT: Do Not Cite or Quote

-------
Table 11-5 (Continued): Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects
Study Site Study Species Reference
Rich fen and 60 kg N/ha/yr as
ombrotrophic urea
'309	S dep = not
reported
N dep = not
reported
In Chamaedaphne
calyculata, fertilization
increased productivity by
87% and biomass by
105%.
In Calamagrostis
canadensis, biomass
increased 600% over
unamended fen plots.
The entire plant
community responded to
N addition with increases
in productivity (82%) and
biomass (160%).
Gogebic	Plant community Iversen et al.
County, Ml consisting of (2010)
Sphagnum spp.,
ericaceous
shrubs, dominant
vascular plants
Carex
oligosperma and
Chamaedaphne
calyculata
Rich fen Addition: two
aqueous
treatments:
groundwater
(0.18 mg N/L,
0.02 mg P/L, 90 mg
Ca2+/L,
pH = 8.0-8.6) and
rainwater
(0.58-0.98 mg N/L,
0.03 mg P/L, pH
6.5-7.0)
Three N addition
levels dissolved in
each aqueous
treatment: no
additional N (low), 1
mg N/L added
(medium), 3 mg N/L
added (high)
Ambient
deposition = not
specified, but
rainwater N is
220-440% higher
than groundwater N
The daily relative growth
rate (RGR) of all four
species decreases 18%
under rainwater rather
than under groundwater.
The switch to rainwater
accounted for 10% of
RGR variation in C.
cuspidata, 7% of RGR
variation in H. vernicosus,
and 32% of RGR
variation in P. squarrosa.
RGR of all four species
decreased by 12% under
medium N addition, and
decreased 24% under
high N addition. N
addition accounted for
4.6% of RGR variation in
B.	pseudotriquetrum,
6.7% of RGR variation in
C.	cuspidata, 18% of
RGR variation in H.
vernicosus, and 23% of
RGR variation in P.
squarrosa.
Mesocosms
constructed
from sand,
peat, and chalk,
with all four
species added
to each
mesocosm,
fully factorial
design: two
aqueous
treatments *
addition
levels x 3P
addition levels
3N
Two bryophyte Andersen et
species common al. (2016)
in rich fens in
Denmark:
Calliergonella
cuspidata and
Bryum
pseudotriquetrum,
and two
bryophyte species
that have been
declining in
abundance over
the past 100 yr:
Hamatocaulis
vernicosus and
Paludella
squarrosa
June 2018
11-47
DRAFT: Do Not Cite or Quote

-------
Table 11-5 (Continued): Nitrogen loading effects upon production and biomass.
Type of Additions or Load Biological and Chemical
Ecosystem (kg N/ha/yr)	Effects
Study Site Study Species Reference
Ombrotrophic
bog
Ambient
deposition = 3.4-5.0
kg N/ha/yr across
the region, divided
for this study into
the following
ranges: 3.4-3.6,
3.8-4.1,	4.4-4.9,
4.9-5.0	kg N/ha/yr
Pitcher plant growth
increased with increasing
N deposition between 3.4
and 4.1 kg N/ha/yr. Plant
growth measured as plant
diameter, which
correlates with pitcher
opening width, pitcher
length, keel width, and
total pitcher width. All
traits positively correlated
with increasing N
deposition between 3.4
and 4.1 kg N/ha/yr.
Deposition above 4.1 kg
N/ha/yr decreased pitcher
plant growth by 8-21% as
measured by plant
diameter (compared to
plant diameter measured
between 3.8. and 4.1 kg
N/ha/yr). Pitcher plant
growth was lower at
higher deposition levels of
4.4-5.0.
Adirondack
Mountains, NY,
(11 sites)
Sarracenia
purpurea
Crumley et al.
(2016)
Calcareous,
rich fen
Addition = 50 kg
N/m2/yr as NO3" or
50 kg N/m2/yr as
NH4+.
Ambient
deposition = 7-10
kg N/m2/yr, based
on Aherne and
Farrell (2002)
NHx decreased
Scorpidium living biomass
67%.
Scragh Bog,
central Ireland
Scorpidium
revolvens and
Sphagnum
con tort urn
Paulissen et
al. (2016)
AG = aboveground; ANPP = aboveground net primary productivity; C02 = carbon dioxide; dep = deposition; ha = hectare;
kg = kilogram; N = nitrogen; NaN03 = sodium nitrate; NH4CI = ammonium chloride; NH4NO3 = ammonium nitrate;
(NH4)2S04 = ammonium sulfate; S = sulfur; yr = year.
11.5. Plant Stoichiometry and Physiology
1	N addition alters the stoichiometry of plant tissue. When inorganic N stores increase in
2	the soil matrix due to N deposition, one of the first plant responses may be to increase N
3	uptake and storage. As a result, the concentration of N in plant tissue (often reported as
4	[N] or %N) is one of the earliest indicators of changing bioavailability of N within an
5	ecosystem. Plants may store the additional N in proteins, use it for further growth if other
6	nutrients and abiotic conditions are not limiting, or curtail growth of belowground,
June 2018	11-48	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
nutrient-acquiring roots. Plant tissue [N] also has important implications for consumers,
as many insect and vertebrate consumers preferentially graze on high N tissues, and plant
[N] can also affect decomposition rates. In systems in which N availability increases
while other nutrient cycles are unaltered, such as in nutrient-poor bogs, nutrient
imbalances develop and plant [K] and [P] decline. This section will consider the effects
of N deposition upon plant tissue [N] as a sensitive indicator of plant response, as well as
the effects of N deposition upon other elements in plant tissue, which would indicate
more detrimental effects of increased N availability upon plant growth.
Plant stoichiometry theory considers the balance of multiple chemical elements in living
tissues. The stoichiometry of plant tissue is often connected to its physiological function.
Most new studies on physiology and stoichiometry are for bogs and freshwater marshes.
There are a few studies on salt water marshes and fens. The new literature is summarized
in the following sections.
11.5.1. Salt Marsh
There are several studies on the effects of N addition on salt marsh plant stoichiometry.
In three California salt marshes, N addition at seven different levels (0, 100, 200, 400,
800, 1,600, and 3,200 kg N/ha/yr) showed that the succulent forb Salicornia depressa
increased leaf %N in a saturating response, with no statistically significant increases in
response above 800 kg N/ha/yr (Vivanco et al.. 2015). There are several N addition
studies at addition levels greater than 500 kg N/ha/yr [(Baldwin. 2013; Nelson and
Zavaleta. 2012; Ryan and Bover. 2012); see Table 11-61.
11.5.2. Mangrove
The N addition levels in all new studies on mangroves are generally higher (greater than
500 kg N/ha/yr) than would be relevant to the effects of N deposition. As perennial
plants, mangroves respond to increasing N loads by altering tissue storage of N as well as
by increasing productivity (Appendix 11.4.3). In salt-marsh mangrove ecotone, Avicennia
germinans seedlings responded to addition of 1,400 kg N/ha with 62% higher leaf tissue
%N, as well as increased N in belowground biomass (Simpson et al.. 2013). In more
mature mangrove ecosystems receiving higher N loads (11,200 kg N/yr per tree), N
addition decreased the efficiency with which Rhizophora mangle resorbed N from
senescing leaves by 7%, which resulted in 39% higher %N in senesced leaves than in
unfertilized trees (Feller etal.. 2009).
June 2018
11-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
11.5.3. Freshwater Marsh
N addition is documented to alter plant stoichiometry and physiology in freshwater
marshes. This topic was not addressed in the 2008 ISA. Five new studies have been
published since 2008.
In freshwater marsh mesocosms planted with wetland obligate graminoid Bolboschoenus
maritimus [cosmopolitan bulrush, multiple synonyms including Schoenoplectus
maritimus, listed as Special Concern by Connecticut and Rhode Island, and as
endangered by Illinois, New Jersey, and New York; (USDA. 2015b)l N addition
increased N stored in plant tissues. Aboveground tissue %N increased 31-114% of %N
in control plants with increasing N loads (32, 65, 108 kg N/ha/yr) added in fall 2009, and
increased 33-67% with increasing N loads (59, 119, 198 kg N/ha/yr) added in the
summer and fall of 2010 (Duguma and Walton. 2014). N addition also increased tissue N
in belowground tissues. In fall 2009, addition of 32 kg N/ha/yr did not affect BG %N, but
the addition of 65 kg N/ha/yr increased BG %N by 26%, and the addition of 108 kg
N/ha/yr increased BG %N by 126 to 1.29% tissue N. In summer and fall 2010, the
addition of 119 and 198 kg N/ha/yr increased BG %N 47 and 37%, with a tissue N value
of 1.34 and 1.25% N, respectively (Duguma and Walton. 2014).
In tidal freshwater marshes of the Tchefuncte River, LA, experimentally added N loads
of 50, 200, 1,200 kg N/ha/yr did not alter the relative dominance of Sagittaria lancifolia,
but did alter N storage in plant tissues. Plant tissue N (mg N/g leaf) increased 0.02 mg for
every 10 kg N/ha/yr of N load. The change in plant N shifted the N:P ratio of plant tissue,
0.04 N: 1 P increase for every additional 10 kg N/ha/yr (Graham and Mendelssohn. 2010).
N addition also decreased the efficiency with which plants resorbed nutrients from
senescing leaves for use in future growth. N addition decreased N resorption efficiency
(%) by 0.1 % for every additional 10 kg N/ha/yr, and decreased P resorption efficiency
(%) by 0.09% for every additional 10 kg N/ha/yr (Graham and Mendelssohn. 2010).
In freshwater marsh mesocosms at the Smithsonian Environmental Research Center
(SERC), native and introduced haplotypes of graminoid Phragmites australis were grown
in a nitrogen addition experiment. The species P. australis includes haplotypes (genetic
lineages) native to North America, as well as invasive haplotypes capable of forming
monocultures that exclude native plant species. Nitrogen fertilization of 250 kg N/ha/yr
increased the construction cost (grams of glucose invested in synthesizing a gram of
biomass) of leaves for the native F haplotype oiP. australis by 5% (Caplan et al.. 2014).
Nitrogen addition did not significantly alter construction costs of the invasive M
haplotype [listed as a noxious weed, banned invasive, or plant pest by Alabama,
Connecticut, Massachusetts, South Carolina, Vermont, and Washington; (USDA. 2015b;
Caplan et al.. 2014)1. These results suggest that N enrichment favors the invasive
June 2018
11-50
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
haplotype over the native haplotype and may facilitate invasion; this is of concern
because P. australis tolerates brackish conditions and invades North American wetlands
all along the freshwater-freshwater tidal-salt marsh continuum.
In tidal freshwater marshes in the Altamaha River Estuary, GA, N addition of 500 kg
N/ha/yr in the 2nd year stimulated productivity (Appendix 11.4.2) to the extent of
diluting N and phosphorus (P) in plant tissues, decreasing leaf N (%) by 22%, and
decreasing leaf P (jxg/g) by 25% (Frost et al.. 2009). In the later years of the experiment,
N addition increased leaf N (%) above control plot leafN by 25-43% (Ketetal.. 2011).
and this increase altered leaf nutrient ratios, decreasing leaf C:N 18-28% and increasing
leaf N:P 25-61% over leaves from control plots. The increases in productivity combined
with alterations in leaf tissue nutrients altered total nutrient pools stored in aboveground
biomass, increasing total aboveground N by 236% and total aboveground P by 169%
(Ket et al.. 2011).
11.5.4. Riparian Wetland
In developing riparian forests at Bonanza Forest LTER, Alaska, N addition of 100 kg
N/ha/yr increased leafN 10% in tree species Alnus incana ssp. tenuifolia growing in late
successional riparian forest, but not in recently colonized or midsuccessional sand bars
(Ruess et al.. 2013). N addition also altered A incana ssp. tenuifolia's internal P cycling
efficiency, decreasing resorption of P from senescing leaves by 21% in midsuccessional
forest (Ruess et al.. 2013).
11.5.5. Bog and Fen
Carnivorous plants grow optimally under low-N conditions, and many are limited in their
distribution to bogs and fens. Recent research suggests that N deposition can disrupt their
N-acquisition strategy of digesting the invertebrates they trap. Purple pitcher plants
(.Sarracenia purpurea) were measured and sampled across a narrow N deposition
gradient (3.4-5.0 kg N/ha/yr) across bogs in the Adirondack Mountains (Crumley et al..
2016). Plant architecture, plant tissue N, and stable isotope N data suggested a growth
and carnivory optima for plants in bogs receiving 3.8-4.1 kg N/ha/yr. At higher
deposition rates of 4.4-4.9 kg N/ha/yr, there was no change in plant tissue N, but there
were negative effects upon growth (see Appendix 11.4.4) and a 40% decrease in
carnivory (Crumley et al.. 2016). The wetland obligate pitcher plants have previously
been shown to experience detrimental impacts of N deposition at physiological and
population levels lYGotelli and Ellison. 2002) in 2008 ISA], Bott et al. (2008) reciprocally
June 2018
11-51
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
transplanted Sarracenia purpurea ssp. purpurea [listed as exploitably vulnerable in New
York, threatened in New York, and endangered in Georgia and Illinois; (USDA. 2015b)!
between ombrotrophic Sapa bog and the rich Cedarburg fen, both in Wisconsin. Leaf %N
was positively and linearly correlated with surface water NOs concentration at the site
where the plants were transplanted (leaf %N = 0.9581 + 0.1667 x |iM NO3 ). confirming
Gotelli's designation of pitcher plants as sensitive indicators ofN deposition (Bott ct al..
2008).
A deposition-induced shift away from insectivory can lead to deficiency in other nutrients
for carnivorous plants. Across a natural deposition gradient in Sweden, the carnivorous
plant Drosera rotundifolia experienced significant declines of insectivory-derived N at
3.81 and 11.30 kg N/ha/yr (Millctt et al.. 2012). Across a broader deposition gradient of
European bogs (0.5-27.0 kg N/ha/yr), D. rotundifolia insectivory declined in a linear
relationship with increasing N deposition, and plants experienced increasing P limitation
as evidenced by increasing N:P tissue concentrations (Millett et al.. 2015). These studies
support previous research showing negative effects of N deposition upon D. rotundifolia
populations (see Appendix 11.7).
Research conducted in European fens and reviewed in the 2008 ISA has shown that N
addition favors the growth of grass and sedge species over peat-forming moss species
(U.S. EPA. 2008a). New research suggests that North American sedge and grass species
experience positive effects of N loading, and that their responses can alter N dynamics of
North American bogs and fens. In a bog in Gogebic County, MI, where sedge Carex
oligosperma [listed as Special Concern in Connecticut, threatened in Ohio and
Pennsylvania, and endangered in Illinois, Massachusetts, and North Carolina; (USDA.
2015b)] was dominant in the plant community, the obligate wetland species responded to
60 kg N/ha/yr addition by decreasing N use efficiency by 89% and N response efficiency
by 84% (Iversen et al.. 2010). The response of the sedge altered ecosystem responses in
the bog, increasing plant community N uptake by 125% over uptake in unamended plots
(Iversen et al.. 2010). These changes increased N stored in biomass pools 171% above
biomass N in control plots and increased the concentration of N in vascular plant tissues
by 19% (Iversen et al.. 2010). Similarly, the grass Calamagrostis canadensis [classified
as wetland facultative or obligate in different U.S. regions by (USDA. 2015b)l. which
was dominant in a closely adjacent rich fen, responded to N addition of 60 kg N/ha/yr by
increasing N uptake. Along with increased productivity (Appendix 11.4.5). this response
expanded the total pool of N stored in biomass by 300% (Iversen et al.. 2010). A similar
mesocosm study of bogs in western Europe across a deposition gradient of 2 to 47 kg
N/ha/yr found that at the highest deposition level, plant tissue N of grasses and sedges
was 120% higher than grasses and sedges at the lowest deposition level (Zaiac and
Blodau. 2016).
11-52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
In the same set of experiments in Gogebic County, MI, woody plants increased their
productivity in response to N loading, but their efficiency in using N declined, suggesting
a mechanism by which N loading lowers N retention of bogs and fens
(Appendix 11.3.1.5). The wetland obligate shrub Chamaedaphne calyculata (threatened
in Illinois and Maryland) responded to 60 kg N/ha/yr addition by increasing plant N
uptake by 75%, which increased productivity (Appendix 11.4) and increased the pool of
N stored in aboveground biomass by 100%. However, nitrogen addition decreased the
efficiency of photosynthesis and metabolism in C. calyculata, as N use efficiency was
81% lower and N response efficiency was 91% lower in fertilized plots (Iversen et al..
2010). In an intermediate fen, tree species Alnns incana ssp. rugosa, speckled alder
[hereafter referred to as A. rugosa, endangered in Illinois; (USDA. 2015b)l. responded to
an added 60 kg N/ha/yr with a 45% decline in N response efficiency [net primary
productivity relative to available soil N; (Iversen et al.. 2010)1. A European study of
mesocosms collected across a deposition gradient of bogs found that 12 kg N/ha/yr
decreased N retention of shrub species, and 47 kg N/ha/yr increased ecosystem total N
stored in shrub biomass (Zaiac and Blodau. 2016).
A long-term fertilization experiment at the Mer Bleue Bog in Ontario, Canada, has
documented the effect of more than a decade of elevated N inputs on the same species in
an ombrotrophic peat bog. Chamaedaphne calyculata responded to 32 g N/ha/yr by
increasing P resorption by 42%, and responded to 64 kg N/ha/yr by increasing P
resorption by 33% (Wang et al.. 2014b). N Plots that had received 64 kg N/ha/yr for
6 years decreased leaf calcium 22% (Wang et al.. 2014b). and leaf Ca was 34% lower in
C. calyculata from plots that received 9 years of 16 kg N/ha/yr than in leaves from
control plots (Bubier etal.. 2011). An observational study across a deposition gradient in
the Adirondack Mountains found effects on C. calyculata stoichiometry at much lower
deposition levels. There was a 27% increase in plant tissue N in C. calyculata when
deposition increased from a range of 3.4-3.6 kg N/ha/yr to 4.9-5.0 kg N/ha/yr (Crumley
etal.. 2016).
After 9 years of N addition at Mer Bleue, leaf C:N ratios were 15% lower in C.
calyculata leaves that received 16 kg N/ha/yr than in leaves from control plots, with
dependent increases in leaf alanine and y-aminobutyric acid (GABA; see Table 11-6 for
equations), amino acid pools that indicate N saturation in the plant (Bubier et al.. 2011).
Leaf aluminum concentrations were 44% lower in these leaves than in leaves from
control plots (Bubier etal.. 2011). In plots that received 64 kg N/ha/yr, there was no
change in leaf C:N, but there were increases in proteins indicative of increased plant N
uptake. C. calyculata leaf total chlorophyll was 84% higher than in control leaves, and
there were 115% increases in leaf alanine and 94% increases in leaf GABA, indicating N
saturation. Leaf manganese was 45% lower in these leaves than in control plot C.
11-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
calyculata, indicating that shrub productivity may be limited by micronutrients in a
N-fertilized bog (Bubier et al.. 2011).
At Mer Bleue, the wetland obligate shrub Rhododendron groenlandicum (synonym:
Ledum groenlandicum, rare in Pennsylvania, threatened in Connecticut, and endangered
in Ohio) responded to 4 years of 64 kg N/ha/yr with a 16% increase in leaf N and a
dependent increase in leaf glutamic acid (Bubicr etal.. 2011). Leaf P declined by 54% of
control leaf P after 4 years of 64 kg N/ha/yr (Bubicr etal.. 2011). After 7 years of 64 kg
N/ha/yr, R. groenlandicum responded with a 23% decrease in leaf magnesium (Mg) and a
186% decrease in Mg resorption from senescing leaves (Wang et al.. 2014b). indicating
increasing limitation by magnesium, which could negatively affect photosynthesis and
enzyme activity within the plant. After 7 years of 32 and 64 kg N/ha/yr addition, R.
groenlandicum increased P and K resorption from senescing leaves; P resorption
increased 12-12.2 times the rate of resorption in leaves from unamended plots, and
increased K resorption by 92-123% (Wang etal.. 2014b).
In the same experiment, after 7 years, N addition shifted the seasonality of gross
photosynthesis, with addition of 64 kg N/ha/yr reducing C uptake 29-45% of control
rates in early summer (May-July), and increasing C uptake 25% above rates in control
plots in September (Larmola et al.. 2013). After 9 years of 16 kg N/ha/yr addition, R.
groenlandicum exhibited altered leaf physiology, doubling maximum carboxylation
capacity (Vcmax) on a per-leaf-mass basis (Bubier et al.. 2011). There was no significant
effect of 4 years of 64 kg N/ha/yr upon VCmax measured during the same growing season
(2008). The study authors posited that plants responded to this level of N availability with
a stress response, allocating N to storage in glutamic acid (nmol glutamic acid/g
leaf = 37.914 + 252.2 x leaf %N) instead of using it in photosynthetic enzymes (Bubier et
al.. 2011).
At Mer Bleue, the wetland facultative shrub Vaccinium myrtilloides (designated sensitive
in Washington, special concern in Connecticut, threatened in Ohio and Iowa, and
endangered in Indiana) responded to nitrogen addition with declines in other leaf
nutrients. Leaf calcium declined 23% in response to 9 years of 16 kg N/ha/yr and
declined 56% in response to 4 years of 64 kg N/ha/yr (Bubier et al.. 2011). indicating that
increasing N will increasingly limit leaf Ca and by extension decrease signaling and
structural integrity of plant tissues. Leaf P decreased by 55% after 9 years of 16 kg
N/ha/yr and by 68% after 4 years of 64 kg N/ha/yr. Leaf manganese (Mn) decreased 57%
under 64 kg N/ha/yr (Bubier et al.. 2011). which could affect photosynthesis or N
assimilation because Mn is a component of enzymes involved in these processes. V.
myrtilloides leaf potassium also declined: 47% in response to 9 years of 16 kg N/ha/yr
and 34% in response to 4 years of 64 kg N/ha/yr (Bubier et al.. 2011). Leaf potassium
11-54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
regulates stomatal movement, activates enzymes, and balances electrical charges
associated with ATP production, so K deficiency can decrease photosynthesis and slow
growth. Leaf aluminum decreased 45% of control leaf aluminum concentrations in plots
receiving 64 kg N/ha/yr for 4 years (Bubier et al.. 2011).
N addition can alter tissue nutrients in nonvascular plants as well. In the Mer Bleue bog,
N addition of 32 kg N/ha/yr decreased moss [Ca] by 34%; N addition of 64 kg N/ha/yr
increased moss [P] by 39%, and decreased moss [Ca] by 42% (Wang et al.. 2016a).
These results are consistent with N addition effects of Ca deficiencies in vascular plants
(see above) and ecosystem P limitation (see Appendix 11.3.2.1.3) within this bog
experiment. Surveys of two European bogs located in different mountain ranges found
that N deposition altered the stoichiometry of Sphagnum moss species. Sphagnum tissue
K concentrations were lower in samples collected from the high N deposition (20-25 kg
N/ha/yr) Jizera Mountains, 64% lower in S. rubrum and 62% lower in S. magellanicum,
than in samples from the lower deposition (12.5 kg N/ha/yr) Jeseniky mountain range. S.
magellanicum [Mg] was also 45% lower at the high N deposition site than at the lower N
deposition site (Jirousek et al.. 2015). Fritz etal. (2014) performed a peat mesocosm
study to test how a history of atmospheric N deposition alters the N uptake rates of
Sphagnum magellanicum, a widely distributed peat-forming species. The researchers
collected blocks of living S. magellanicum from a pristine bog in Argentina, which
received an annual load of 1-2 kg N/ha/yr, and also from a polluted bog in the
Netherlands, which received an annual load of 20-30 kg N/ha/yr. Under controlled
conditions, nitrogen solutions of 1, 10, or 100 |imol N/L were added to the block. Rates
of N uptake were significantly different among bog samples (Argentine or Dutch) for the
10 or 100 |imol solutions, with the pristine (Argentine) Sphagnum absorbing nitrogen at
1.4-2.6 times the rate of the uptake rate of the Dutch bog, which receives a higher annual
N load (Fritz et al.. 2014). A mesocosm study of four European moss species found that
applying rainwater (elevated N and P) rather than groundwater increases moss tissue [N]
and [P], while applying additional N in controlled treatments decreases moss tissue [P]
(Andersen et al.. 2016). A different European mesocosm study found that Sphagnum spp.
plant tissue N increased across a N deposition gradient of 2 to 47 kg N/ha/yr, such that
C:N ratio declined as a linear function of the natural log of increasing N deposition
(Zaiac and Blodau. 2016).
Recent research has addressed differential responses of bogs and fens to reduced or
oxidized forms of N addition. In an Irish fen dominated by Sphagnum contortum and
brown moss Scorpidium revolvens, 50 kg N/ha/yr was applied as NOs, in one treatment
and NH4+ in another treatment (Paulissen et al.. 2016). Reduced N forms elicited stronger
responses in tissue chemistry for each moss species. In S. contortum, oxidized N
increased tissue N 29%, tissue P 28%, and concentrations of alanine (a free amino acid
11-55
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
hypothesized to be a mechanism of N storage) 31%; an equivalent addition of reduced N
decreased tissue N 13%, tissue K 49%, and dramatically increased (by half to 66 times
the control) the concentrations of five free amino acids storing excess N in plant tissue. In
the more sensitive species S. revolvens, reduced N increased tissue N 76% but decreased
tissue K 53%, and increased concentrations of arginine (a free amino acid) 7.3 times
concentrations in control plots; these same metrics were unaffected by the equivalent
oxidized N treatment (Paulisscn et al.. 2016). The stronger effects of reduced N on plant
chemistry in S. contortum than S. revolvens may represent species-level differences in N
tolerance, as S.contortum biomass did not change in response to N addition (see
Appendix 11.4.4).
The Whim Bog experiment in Scotland added N as either ammonium or nitrate to
simulate wet deposition, which with ambient N deposition of 8 kg N/ha/yr resulted in
treatment N loads of 16, 32, and 64 kg N/ha/yr. As in other wetlands, the bog had
divergent responses to oxidized and reduced N addition. Sphagnum capillifolium plant
tissue N and N:P increased in response to increasing oxidized N in a saturating function,
but responded to increasing reduced N in a linear function (Chiwa et al.. 2016). Reduced
N addition also resulted in increasing cation concentrations in pore water, suggesting
cation leaching by the moss species as a result of N uptake. These results indicate a
stronger effect of reduced N upon Sphagnum growth or P limitation, and these
physiological changes also affected N cycling and water quality within the bog (see
Appendix 11.3.1.5).
11.5.6. Summary Table
Table 11-6 Nitrogen loading effects upon plant stoichiometry and physiology.
Type of Additions or Load Biological and
Ecosystem (kg N/ha/yr)	Chemical Effects Study Site Study Species	Reference
Coastal salt
marsh
Addition: 100, 200,
400, 800, 1,600,
3,200 kg N/ha/yr as
urea
N dep = 3-5 kg
N/ha/yr as reported
in Tonnesen et al.
(2007)
Salicornia leaf %N
increases in a
saturating response to
added N (Nadd, as g
N/m2/yr), leaf
N = -0.73
Nadd + 2.99.
x e
Morro Bay
National
Estuary,
Carpinteria
Salt Marsh
Reserve,
Tijuana River
Reserve
Estuary, CA
Salicornia
depressa
(Salicornia
virginica) stands
Vivanco et al.
(2015)
June 2018
11-56
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Coastal salt
marsh
3,000 kg N/ha/yr as
NH3NO3
N dep = not
reported
S dep = not
reported
Plant [N] increased by
224 and 33%, and N
pools in AG vegetation
increased by 60 and
84% in successive
summers.
Elkhorn
Slough,
Monterey Bay,
CA
Sarcocornia
pacifica
dominant,
Distichlis spicata,
Frankenia salina,
and Jaumea
carnosa
Nelson and
Zavaleta (2012)
Estuarine salt
marsh
1,337 kg N/ha/yr as In S. pacifica, tissue N China Camp Sarcocornia
urea
N dep = not
reported
S dep = not
reported
increased 28-46%. D.
spicata tissue N
increased 19%. J.
carnosa tissue N
increased by 54%.
State Park, CA
pacifica
dominant,
Distichlis spicata,
and Jaumea
Ryan and Bover
(2012)
carnosa
Mangrove/
salt marsh
ecotone
1,400 kg N/ha/yr
N dep = not
reported
S dep = not
reported
Mangrove leaf
production increased
by 42%. Leaf tissue
%N increased 62%.
Root %N increased
9% and root C:N
decreased 25%.
Merritt Island
National
Wildlife
Refuge, FL
Avicennia
germinans
Simpson et al.
(2013)
Mangroves
No areal
reported;
N/yr per
N dep =
reported
S dep =
reported
rate
; 11,200 kg
tree
not
not
Fertilization increases
growth rate (shoot
elongation) by 290% in
the fringe zone and by
1,340% in the scrub
zone. In the fringe
zone, fertilization
decreases resorption
efficiency of N from
dying leaves by 7%,
resulting in senesced
leaves containing 39%
more N than control
senesced leaves.
Indian River
Lagoon, FL
Rhizophora
mangle
Feller etal. (2009)
Freshwater
marsh
Fall 2009: 32 (low),
65 (medium), 108
(high) kg N/ha/yr as
(NH^SCM; summer
2010: 59 (low), 119
(medium), or 198
(high) kg N/ha/yr as
(NH4)2S04
N dep = not
reported
S dep = not
reported
N addition increased
S. maritimus %N of
aboveground tissue
31-114% in 2009 and
33-67% in 2010. N
addition (medium,
high) increased S.
maritimus %N of
belowground tissue
26-126% in 2009 and
37-47% in 2010.
Mesocosms at
UC Riverside
research
station, CA
Schoenoplectus
maritimus, Culex
tarsalis, and
Anopheles
hermsi
Duquma and
Walton (2014)
June 2018
11-57
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Freshwater
tidal marsh
(growth
chamber)
250 g N/ha/yr as
NH4CI solution
N dep = not
reported
S dep = not
reported
Nitrogen increases
construction cost of
leaves by 5% in native
Phragmites. Nitrogen
under elevated [CO2]
increases construction
costs by 6% in
invasive Phragmites.
Mesocosms at
Smithsonian
Environmental
Research
Center, MD
Phragmites
australis (native F
haplotype and
invasive M
haplotype)
Caplan et al. (2014)
Freshwater
tidal marsh
500 kg N/ha/yr as
NH4CI or urea
N addition decreased Altamaha Zizaniopsis
leaf [N] by 99% and River, GA miliacea
leaf [P] by 25%.
Frost et al. (2009)
Freshwater
marsh
500 kg N/ha/yr as
NH4CI or urea
N dep = not
reported
S dep = not
reported
In Z. miliacea, leaf %N
increased 25-43% so
that leaf C:N
decreased 18-28%
and leaf N:P increased
25-61%. Total N and
total P in aboveground
pools increased 236
and 169%,
respectively.
Altamaha
Estuary, GA
Zizaniopsis
miliacea,
Pontederia
cordata, and
Sagittaria
lancifolia
Ket et al. (2011)
Freshwater
tidal marsh
50, 200, or 1,200 kg
N/ha/yr as
Nutralene
methylene urea
N dep = not
reported
S dep = not
reported
In S. lancifolia, N
addition (Nadd as kg
N/ha/yr) increased
plant tissue [N],
[N] = 0.00226 x Nadd +
23.271; and N:P ratio,
N:P = 0.00404 x
Nadd + 16.853.
In S. lancifolia,
resorption efficiencies
(RE) declined with
increasing N addition,
for nitrogen:
NRE = -0.01048 xNadd
+ 31.796, for
phosphorus:
PRE = -0.00915 xNadd
+ 62.778.
Tchefuncte
River,
Madisonville,
LA
Oligohaline plant
community
dominated by
Sagittaria
lancifolia,
Eleocharis fallax,
and Polygonum
punctatum
Graham and
Mendelssohn
(2010)
Freshwater
Addition: 670 kg
Foliar N increased in
Nanticoke Plant community Baldwin (2013)
tidal marsh
N/ha/yr
A. calamus (by 23%)
River, MD and

S dep = not
and Typha spp. (by
DE

reported
47%).


N dep = not



reported


June 2018
11-58
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Riparian
floodplain
successional
forest
100 kg N/ha/yr
N dep = not
reported
S dep = not
reported
Specific leaf mass
decreased by 12 and
11% in early and late
successional forest.
Leaf N increased 10%
in late successional
forest. Leaf P
resorption decreased
21%.
Bonanza
Forest LTER,
AK
Alnus incana ssp.
tenuifolia and
associated
Frankia strains
Ruess et al. (2013)
Ombrotrophic
bog
Ambient deposition
= 3.4-5.0 kg
N/ha/yr across the
region, divided for
this study into the
following ranges:
3.4-3.6, 3.8-4.1,
4.4-4.9, 4.9-5.0 kg
N/ha/yr
Pitcher plant foliar N
content increased 78%
between low N
deposition (3.4-3.6 kg
N) and medium N
deposition (3.8-4.1 kg
N). There was no
difference in foliar N
between medium and
higher deposition
rates.
Stable isotope (815N)
and mass balance
suggest that 68% of
pitcher plant N was
derived from carnivory
under low N
(3.4-3.6 kg N), 92%
under medium N
(3.8-4.1 kg N), and
55% under higher N
(4.4-4.9 kg N).
Increase from medium
to higher N deposition
decreased pitcher
plant carnivory by
40%.
Chamaedaphne
calyculata foliar N
increased 27% from
the lowest (3.4-3.6 kg
N) to the highest
deposition (4.9-5.0 kg
N) sites.
Adirondack
Mountains, NY
(11 sites)
Sarracenia
purpurea
Crumley et al.
(2016)
Bog
N deposition
gradient: 1.94,
and 11.30 kg
N/ha/yr
N from insectivory
3.81, decreases with
deposition: 55% from
insects under 1.94 kg
N/ha/yr, 20-30%
under 3.81 and
11.30 kg N/ha/yr.
Three bogs in
Sweden
Drosera
rotundifolia
Millett et al. (2012)
June 2018
11-59
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of Additions or Load Biological and
Ecosystem (kg N/ha/yr)	Chemical Effects Study Site Study Species	Reference
Ombrotrophic N deposition	N derived from	13 bogs across Drosera	Millett et al. (2015)
bog	gradient:	insectivory decreases Europe	rotundifolia
0.5-27.0 kg N/ha/yr with increasing N
deposition (Ndep as g
N/m2/yr) in a linear
relationship:
815N = -2.090 x Ndep
- 0.199, R2 = 0.43.
Plant tissue N:P
increases with
increasing N
deposition in a
logarithmic
relationship: tissue
N:P = 1.29 x In(Ndep)
+ 10.00, R2 = 0.62.
June 2018
11-60
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Ombrotrophic
peat bog
16 (low), 32
(medium), or 64
(high) kg N/ha/yr as
N (NH4NO3) or NPK
(NH4NO3 and
KH2PO4)
N dep = 8 kg
N/ha/yr
S dep = not
reported
In R. groenlandicum,
leaf%N increased by
25% (low N) and 16%
(high N), and leaf C:N
ratios decreased by
19% (low N) and 15%
(high N). Leaf P
declined by 54% (high
N). In low N addition,
R. groenlandicum
Vcmax increased by
76% on a per-area
basis or doubled on a
per-mass basis. In C.
calyculata, leafC:N
ratios decreased by
15% (low N) and 33%
(high N), and total
chlorophyll increased
by 84% (high N), just
as concentrations of
leaf alanine increased
115% (high N) and
GABA increased 94%
(high N). Leaf Ca
declined 34% (low N)
and 17% (high N), leaf
Mn declined 45% (high
N), and leaf Al
declined 44% (low N).
In V. myrtilloides, leaf
Ca declined 23% (low
N) and 56% (high N),
leaf P declined by 55%
(low N) and 68% (high
N), leaf Mn declined
57%, leaf K declined
47% (low N) and 34%
(high N), and leaf Al
declined 45% (high N).
Amino acid leaf
concentrations (y, as
nmol/g) increased with
leaf N (x, as %N): in R.
groenlandicum,
glutamic acid
increased
y = 252.2x+ 37.914; in
C. calyculata, alanine
increased
y = 221.07x- 52.931,
and GABA increased
y = 276.82x-98.682.
Mer Bleue
Bog, Ontario,
Canada
Three ericaceous
shrubs:
Vaccinium
myrtilloides,
Rhododendron
groenlandicum
(formerly Ledum
groenlandicum),
and
Chamaedaphne
calyculata
Bubieretal. (2011)
June 2018
11-61
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Ombrotrophic
peat bog
16 (low), 32
(medium), or 64
(high) kg N/ha/yr as
N (NH4NO3) or NPK
(NH4NO3 and
KH2PO4)
Under high N, gross
PSN declined 29-45%
in May-July, but
increased 25% in
September.
Mer Bleue
Bog, Ontario,
Canada
Shrub species
(Vaccinium
myrtilloides,
Rhododendron
groenlandicum,
Chamaedaphne
calyculata), and
mosses
(Sphagnum
magellanicum,
Sphagnum
capillifolium,
Polytrichum
strictum)
Larmola et al.
(2013)
Ombrotrophic
peat bog
16 (low), 32
(medium), or 64
(high) kg N/ha/yr as
N (NH4NO3) or NPK
(NH4NO3 and
KH2PO4)
N dep = 5-6 kg
N/ha/yr
S dep = not
reported
High N decreased leaf
Ca (mg/cm2) by 22%
in C. calyculata and
decreased leaf Mg
(mg/cm2) by 23% in R.
groenlandicum. In C.
calyculata, medium N
increased P resorption
by 42% as high N
increased it by 33%. In
R. groenlandicum, N
addition (medium N,
high N) increased P
resorption by
12-12.2x, and
increased K resorption
by 92-123%. High N
decreased Mg
resorption 186% in R.
groenlandicum.
Mer Bleue
Bog, Ontario,
Canada
Chamaedaphne
calyculata and
Rhododendron
groenlandicum
(formerly Ledum
groenlandicum)
Wang et al. (2014b)
Rich fen and
No addition
Leaf N (y, as %N) was
Cedarburg fen
Sarracenia
Bott et al. (2008)
ombrotrophic
N dep = not
reported
positively correlated
and Sapa bog,
purpurea ssp.

bog
with surface water
Wl
purpurea


NO3" concentration




S dep = not
(x):
y= 0.1667X+ 0.9581




reported



June 2018
11-62
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of Additions or Load
Ecosystem (kg N/ha/yr)
Biological and
Chemical Effects Study Site
Study Species	Reference
Plant community Iversen et al. (2010)
consisting of
Sphagnum spp.,
ericaceous
shrubs, dominant
vascular plants
Carex
oligosperma and
Chamaedaphne
calyculata
Ombrotrophic
bog
60 kg N/ha/yr as
urea
S dep = not
reported
N dep = not
reported
In Chamaedaphne Gogebic
calyculata, fertilization County, Ml
increases plant N
uptake by 75%, N
stored in biomass by
100%, productivity by
87%.
Fertilization decreases
C. calyculata N use
efficiency by 81% and
N response efficiency
by 91%.
In Carex oligosperma,
fertilization decreases
N use efficiency by
89% and N response
efficiency by 84%.
The entire plant
community responded
to N addition with
increases in N uptake
(125%), N stored in
biomass (171%),
productivity (82%), and
[N] in new tissue
(19%).
Fen
60 g N/ha/yr as
urea
N dep = not
reported
S dep = not
reported
In A. rugosa,	Gogebic
fertilization decreases County, Ml
N response efficiency
by 45%.
Dominant	Iversen et al. (2010)
vascular plants
Calamagrostis
canadensis and
Alnus rugosa
June 2018
11-63
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Ombrotrophic
peatland
Ambient
deposition = 2 kg
N/ha/yr at DS;
12 kg N/ha/yr at
WM; 47 kg N/ha/yr
at FS
Mesocosms were
established with
NH4NO3 solutions
that mimicked
relative N
deposition levels
15N tracer was
added in
48 applications over
6 mo at a rate of
23 kg N/ha/yr
At 12 kg N/ha/yr, 15N
retention efficiency in
shrubs decreased
65% compared to DS.
At 47 kg N/ha/yr, plant
tissue N of graminoids
increased 120%
compared to DS. Total
N stored in shrubs
increased 360%
compared to DS.
Across five sites and
all mesocosms,
Sphagnum C:N was a
function of N
deposition (Ndep, as g
N/m2/yr):
C:N = -22.9 * ln(NdeP)
+ 68.6.
Data from LV and CF
mesocosms are not
considered because
mesocosm N addition
levels (CF: 300%
increase over DS N
solution, LV: 700%
increase over DS) did
not reflect ambient N
deposition differences
(CF: 750% increase
over DS deposition,
LV: 300% increase
over DS).
Across five sites and
all mesocosms,
Sphagnum C:N was a
function of N
deposition:
C:N = -22.9 * ln(NdeP)
+ 68.6.
Mesocosms
Sphagnum
constructed
capillifolium, S.
using peat bog
fallax, S.
cores from
magellanicum, S.
sites in
papillosum, S.
northern and
pulchrum, S.
western
rubellum,
1 Europe—
Andromeda
Degero
polifolia, Calluna
Stormyr,
vulgaris, Erica
Sweden (DS);
tetralix, Rubus
Fenn's,
chamaemorus,
Whixall, and
Vaccinium
Bettisfield
oxycoccos,
Mosses NNR,
Eriophorum
U.K. (WM);
vaginatum,
and
Eriophorum
Frolichshaier
angustifolum
Sattelmoor,

Germany (FS)

(Peat cores

also collected

at Little

Vildmose,

Denmark [LV]

and Cors

Fochno,

Wales, U.K.

[CF])

Zaiac and Blodau
(2016)
June 2018
11-64
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Ombrotrophic
peatland
Addition = 16 kg
N/ha/yr (5N, or 5x
background
deposition), 32 kg
N/ha/yr (1 ON), or
64 kg N/ha/yr
(20N), all as
NH4NO3
Ambient (wet)
deposition = 8 kg
N/ha/yr and up to
0.26 kg P/ha/yr
N addition of 32 kg
N/ha/yr decreased
moss [Ca] by 34%. N
addition of 64 kg
N/ha/yr increased
moss [P] by 39%, and
decreased moss [Ca]
by 42%.
Mer Bleue bog,
Ottawa,
Canada
Evergreen
shrubs:
Chameadaphne
calyculata and
Rhododendron
groenlandicum
dominant; sparse
Kalmia
angustifolia
Deciduous
shrubs: Vaccinium
myrtilloides
Mosses:
Sphagnum
capillifolium, S.
magellanicum,
and Polystrichum
strictum
Wang et al. (2016a)
Ombrotrophic
bog
Wet + dry
deposition,
estimated by
Jirousek et al.
(2011)
Jireza: 20-25 kg
N/ha/yr (high N)
Jeseniky: 12.5 kg
N/ha/yr (low N)
S. rubellum [K] was
64% lower at high N
site. S. magellanicum
[K] was 62% lower and
[Mg] was 45% lower at
high N site.
Two sites:
Sphagnum fallax,
Jizerka in
Sphagnum
Jireza
magellanicum,
Mountains
and Sphagnum
(warm
rubellum/russowii
suboceanic

climate) and

Vozka in

Jeseniky

Mountains

(2°C colder)

Laboratory
Sphagnum
incubation of
magellanicum
moss from

pristine site in

Argentina and

N polluted site

in Netherlands

Jirousek et al.
(2015)
Bog
Laboratory
incubation. Moss
from Argentina
(54.75°S, 68.33°W)
and Netherlands
(52.82°N, 2.42°E)
Pristine site
(Argentina): 1-2 kg
N/ha/yr, N polluted
site (Netherlands):
20-30 kg N/ha/yr
N uptake of 10 and
100 pmol N/L solutions
was 40-160% more
rapid by Sphagnum
from the pristine
(Argentinian) bog than
the Dutch bog.
Fritz et al. (2014)
June 2018
11-65
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Rich fen	Addition: two
aqueous
treatments:
groundwater
(0.18 mg N/L,
0.02 mg P/L, 90 mg
Ca2+/L,
pH = 8.0-8.6) and
rainwater
(0.58-0.98 mg N/L,
0.03 mg P/L, pH
6.5-7.0)
Three N addition
levels dissolved in
each aqueous
treatment: no
additional N (low),
1 mg N/L added
(medium), 3 mg N/L
added (high)
Ambient
deposition = not
specified, but
rainwater N is
220-440% higher
than groundwater N
Groundwater vs.
rainwater: tissue [N]
and tissue [P] were
higher in rainwater
than in groundwater
treatment.
N addition treatment:
tissue [P] was lower in
medium and high N
addition mesocosms.
Mesocosms
constructed
from sand,
peat, and
chalk, with all
four species
added to each
mesocosm,
fully factorial
design: two
aqueous
treatments * 3
N addition
levels x 3 P
addition levels
Two bryophyte
species common
in rich fens in
Denmark:
Calliergonella
cuspidata and
Bryum
pseudotriquetrum
Two bryophyte
species that have
been declining in
abundance over
the past
100 years:
Hamatocaulis
verrucosus and
Paludella
squarrosa
Andersen et al.
(2016)
June 2018
11-66
DRAFT: Do Not Cite or Quote

-------
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of
Ecosystem
Additions or Load
(kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Calcareous,
rich fen
Addition = 50 kg
N/m2/yr as NO3" or
50 kg N/m2/yr as
NH4+
Ambient
deposition = 7-10
kg N/m2/yr, based
Aherne and Farrell
(2002)
In Scorpidium, NHx
increased tissue N
76% and decreased
tissue K 53%. NHx
increased Scorpidium
free arginine
concentration 7.3x
control.
In Scorpidium, NOx
increased
phosphomonoesterase
(PMEase) 60%,
indicating increased P
limitation.
In Sphagnum, NHx
increased PMEase
70%, indicating
increased P limitation.
NHx decreased
Sphagnum tissue N
13% and tissue K
49%. NHx increased
the concentrations of
free amino acids in
Sphagnum: alanine by
55%, arginine 66x
control concentrations,
asparagine 39x
control, glutamine 65*
control, and serine
3.8x control.
In Sphagnum, NOx
increased tissue N
29% and tissue P
28%. NOx increased
Sphagnum free
alanine concentration
31%.
Scragh Bog,
Central Ireland
Scorpidium
revolvens and
Sphagnum
con tort um
Paulissen et al.
(2016)
June 2018
11-67
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
Table 11-6 (Continued): Nitrogen loading effects upon plant stoichiometry and
physiology.
Type of Additions or Load
Ecosystem (kg N/ha/yr)
Ombrotrophic Addition = 8, 24,
peatland	and 56 kg N/ha/yr
as either NH4+ or
NO3" since 2002
Ambient deposition
total N = 8 kg
N/ha/yr, 3 kg
N/ha/yr as wet NOx,
3 kg N/ha/yr as wet
NHx, and 2 kg
N/ha/yr as dry NHx
Biological and
Chemical Effects Study Site
Tissue N in Sphagnum Whim bog,
capitulum, and tissue Edinburgh,
N and N:P in	Scotland
Sphagnum stems,
increased in a
saturating
(i.e., logarithmic)
relationship with
increasing NOx
deposition.
Tissue N and N:P in
Sphagnum capitulum,
and tissue N in
Sphagnum stems,
increased in a linear
relationship with
increasing
NHxdeposition.
Study Species	Reference
Sphagnum moss Chiwa et al. (2016)
(Sphagnum
capillifolium, a
hummock-
forming species)
Heathland
community:
Calluna vulgaris,
Eriophorum
vaginatum,
Hypnum
jutlandicum,
Pleurozium
schreberi, and
Cladonia
portentosa
AG = aboveground; Al = aluminum; ANPP = aboveground net primary productivity; C = carbon; Ca = calcium; cm = centimeter;
C02 = carbon dioxide; dep = deposition; GABA = gamma-Aminobutyric acid; ha = hectare; K = potassium; kg = kilogram;
KH2PO4 = potassium phosphate; L = liter; LTER = Long Term Ecological Research; |jmol = micromole; mg = milligram;
Mn = manganese; N = nitrogen: NH4CI = ammonium chloride; NH4NO3 = ammonium nitrate; (NH4)2S04 = ammonium sulfate;
N03" = nitrate; NPK= nitrogen, phosphorus, potassium; NRE = nitrogen resorption efficiency; P = phosphorus; PRE = phosphorus
resorption efficiency; S = sulfur; yr = year.
11.6. Plant Architecture
Plant architecture (plant height, branching) in wetlands is an important endpoint because
the architecture of the dominant plants determines the availability of nesting habitats or
refugia from flooding and predators. Plant architecture is defined as the
three-dimensional organization of the plant body. For the parts of the plant that are
aboveground, this includes the branching pattern, as well as the size, shape, and position
of leaves and flower organs (Table 11-7). The 2008 ISA did not have sufficient data on N
addition effects on wetland plant architecture to consider plant architecture as a specific
endpoint.
11.6.1. Salt Marsh
There are several new studies on salt marsh plant architecture. Darby and Turner (2008)
reported that for Spartina alterniflora, stem height increased linearly with N load starting
at 230 kg N/ha/yr, although regression equations were not included. Stem density was
June 2018
11-68
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
significantly higher than in unamended plots only for the highest N load, 3,720 kg
N/ha/yr, which increased stem density by 43%. There are two other addition studies with
N addition rates above 500 kg N/ha/yr that evaluated effects in U.S. salt marshes (Ryan
and Bover. 2012; Davevetal.. 2011).
The Plum Island tidal eutrophication experiment added nitrate and phosphate to salt
marsh tidal creeks to achieve nitrate concentrations 15 times the concentrations of
unamended creeks. In the last 2 years of this experiment, there was N but no P
enrichment, and 620 kg N/ha/yr increased Spartina alterniflora plant height 4% in low
marsh (Johnson et al.. 2016a). In the high marsh, 140 kg N/ha/yr decreased Distichlis
spicata height 4% (Johnson et al.. 2016a). Across the course of the N and P addition,
alterations in plant carbon allocation decreased vascular plant root: shoot biomass ratio by
31% (Dccgan et al.. 2012). With the decrease in belowground support and the decreases
in shoot lignin (which provides structural support in plant shoots, see Stoichiometry
section), a 5% increase in shoot height made plants top-heavy. As a result, lodging (plant
stems collapsing to marsh substrate where they are inundated by tides), which was not
observed in control marshes, affected 41% of the marsh area in enriched marsh (Deegan
et al.. 2012).
11.6.2. Mangrove
In Florida mangroves, Whigham et al. (2009) added 100 kg N/ha/yr to plots of dwarf
Avicennia germinans, black mangrove, which increased the number of new branches
produced 150% above control plots.
11.6.3. Freshwater Tidal Marsh
There are several new studies on freshwater tidal wetlands. In freshwater tidal marshes of
Altamaha River Estuary, GA, N addition (500 kg N/ha/yr) increased the number of leaves
per m2 52% and increasing plant height by 25-40% over the course of the 5-year
experiment (Ket et al.. 2011). Duguma and Walton (2014) constructed freshwater
wetland mesocosms at UC Riverside, planting Bolboschoenus maritimus (formerly
Schoenoplectus maritimus) and alkali bulrush, and monitoring plant responses as well as
the use of mesocosms by mosquito species (Appendix 11.8.2). N addition of 32 kg
N/ha/yr increased plant height 23% with no effect upon plant height when 65 or 108 kg
N/ha/yr were added in 2009. However, in 2010 when the experiment was initiated earlier
in the year (5 August start date in 2009, 22 March start date in 2010), all N addition
June 2018
11-69
DRAFT: Do Not Cite or Quote

-------
1
2
levels (59, 119, 198 kg N/ha/yr) increased plant height 28-44% above control plants
(Duguma and Walton. 2014).
11.6.4. Riparian Wetland
In developing riparian forests at Bonanza Forest LTER, AK, N addition of
100 kg N/ha/yr altered leaf structure of the tree species Alnus incana ssp. tenuifolia
(Ruess et al.. 2013). In the mostly recently colonized sand bars, N addition decreased
specific leaf mass (SLM) of A. incana ssp. tenuifolia 12%, and in midsuccessional
riparian forest SLM decreased 11% with N addition.
11.6.5. Summary Table
Table 11-7 Nitrogen loading effects upon architecture.
Type of
Ecosystem
Additions or
Load (kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Coastal salt
marsh
4,200 kg N/ha/yr
as NH4NO3
S dep = not
reported
N dep = not
reported
N addition increased
rhizome diameter by
1% in shallow and 5%
in deep (10-20 cm)
sediments.
Goat Island,
SC
Spartina
alterniflora
Davev et al. (2011)
Coastal salt
marsh
230, 465, 930,
1,860, or 3,720 kg
N/ha/yr as
(NH4)2S04
S dep = not
reported
N dep = not
reported
Stem density and stem
height increased
linearly with N load
(regressions not given).
Stem densities
increased 43% in
response to 3,720 kg N.
LUMCON,
Cocodrie, LA
Spartina
alterniflora
Darby and Turner
(2008)
Estuarine
marsh
N/ha/yr as urea.
S dep = not
reported
N dep = not
reported
In S. pacifica, height
China Camp Sarcocornia
increased 207%,
State Park, CA pacifica
number of branches
dominant (C3
increased 135-283%.
succulent
J. carnosa height
shrub),
increased 504%.
Distichlis

spicata (C4

grass), and

Jaumea

carnosa (C3

semisucculent

forb)
Ryan and Bover
(2012)
June 2018
11-70
DRAFT: Do Not Cite or Quote

-------
Table 11-7 (Continued): Nitrogen loading effects upon architecture.
Type of
Ecosystem
Additions or
Load (kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species
Reference
Salt marsh
Nitrate and
phosphate
dissolved in tidal
inflows to raise
aqueous NO3"
concentrations to
70-100 |jM and
raise PO43" to
5-7 |jM.
Background load
in tides: 5 |jM
NO3-, 1 |JM PO43-
N dep = not
reported
S dep = not
reported
Root:shoot ratio
decreases 31%.
Shoot height increases
5%.
Lodging affects 41% of
area in enriched
marshes, is not
observed in control
marsh.
Plum Island
Estuary, MA
Primary tidal
creeks with
Spartina
alterniflora in
low marsh
along creeks,
and Spartina
patens in high
marsh
platforms
Deeaan et al.
(2012)
Salt marsh
Addition of nitrate
dissolved in
incoming tide
2011-2012,
addition of N at
70-100 |jM
NaNC>3" in tide for
added load of
620-1,200 kg
N/ha/yr in low
marsh and
70-140 kg N/ha/yr
in high marsh
Ambient
deposition = not
specified
Low marsh in enriched
creeks received an
additional N load of
620 kg N/ha/yr in 2012.
That year, N addition
increased shoot height
4% above S.alterniflora
in low marsh at
reference creeks.
High marsh in enriched
creeks received an
additional N load of
140 kg N/ha/yr in 2011.
In D. spicata, N addition
decreased shoot height
4%.
Plum Island
Sound
Estuary, MA
Spartina
alterniflora,
Distichlis
spicata, and
Spartina patens
(high marsh);
Spartina
alterniflora (low
marsh)
Johnson et al.
(2016a)
Mangrove
100 kg N/ha/yr
N addition increased
Indian River
Avicennia
Whiqham et al.


the number of new
Lagoon, FL
germinans and
(2009)


branches 150% in
(impoundment
associated



Avicennia.
SLC-24)
sediments

Freshwater
marsh
Fall 2009: 32
(low), 65
(medium), 108
(high) kg N/ha/yr
as (NH4)2S04;
Summer 2010: 59
(low), 119
(medium), or 198
(high) kg N/ha/yr
as (NH4)2S04
N dep = not
reported
S dep = not
reported
N addition increased S.
maritimus height 23% in
2009 (low N), and
28-44% in 2010.
Mesocosms at
UC Riverside
research
station, CA
Schoenoplectus
maritimus,
Culex tarsalis,
and Anopheles
hermsi
Duauma and
Walton (2014)
June 2018
11-71
DRAFT: Do Not Cite or Quote

-------
Table 11-7 (Continued): Nitrogen loading effects upon architecture.
Type of Additions or	Biological and
Ecosystem Load (kg N/ha/yr) Chemical Effects Study Site Study Species Reference
Tidal
500 kg N/ha/yr as
In Z. miliacea, leaf
Altamaha
Zizaniopsis
Ket et al. (2011)
freshwater
NH4CI or urea
number increased 52%
Estuary, GA
miliacea,

marsh

and plant height
increased 25-40%.

Pontederia
cordata, and
Sagittaria
lancifolia

Riparian
100 kg N/ha/yr
Specific leaf mass
Bonanza
Alnus incana
Ruess et al. (2013)
floodplain
N dep = not
reported
S dep = not
reported
decreased by 12 and
Forest LTER,
ssp. tenuifolia

successional
11% in early and late
AK
and associated

forest
successional forest.

Frankia strains

Ombrotrophic
16 (low), 32
In medium and high
Mer Bleue
Bog plant
Juutinen et al.
peat bog
(medium), or 64
NPK treatments, shrub
Bog, Ontario,
community:
(2010)

(high) kg N/ha/yr
canopies grew 72 and
Canada
dwarf shrub


as N (NH4NO3) or
82% taller than

species and


NPK (NH4NO3 and
controls.

mosses


KH2PO4)
There were no

Sphagnum


S dep = not
significant effects of

magellanicum,


reported
N dep = 8 kg
medium or high N on
shrub canopies.

Sphagnum
capillifolium,
and


N/ha/yr


Polytrichum
strictum

AG = aboveground; ha = hectare; kg = kilogram; KH2P04 = potassium phosphate; LTER = Long Term Ecological Research;
LUMCON = Louisiana Universities Marine Consortium; N = nitrogen; NH4CI = ammonium chloride; NH4N03 = ammonium nitrate;
(NH4)2S04 = ammonium sulfate; P = phosphorus; S = sulfur; yr = year.
June 2018
11-72
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
11.7. Demography
Plant demography is the change in plant population size and structure through time.
Changes to plant demography will determine the long-term stability of wetlands.
Information on the effect of N loading on wetland plant demography, including
establishment, survival, and reproduction is sparse. Evidence from salt marsh ecosystems
shows mixed responses to N loading: two pre-2008 studies found negative effects of N
addition upon reproduction in invasive Spartina foliosa and Spartina hybrids (Tyler et al..
2007). and a positive effect upon reproduction of native Salicornia bigelovii (Bover and
Zedler. 1999). Work reviewed in the 2008 ISA on Sarraceniapurpurea, or northern
pitcher plant, found that increases in N deposition increased population extinction risk
(Gotclli and Ellison. 2006. 2002). which serves as the scientific basis for the wetland
critical load values (Appendix 11.9). Additionally, an older study of Drosera rotundifolia
in a Swedish bog found that N addition levels of 20 and 40 kg N/ha/yr significantly
decreased survivorship in comparison to control plots or N addition of 5 or 10 kg N/ha/yr
(Rcdbo-Torstcnsson. 1994). The sections below describe two new studies that find
negative impacts of high N upon the demography of wetland plants.
11.7.1.	Mangrove
Lovelock et al. (2009) considered a global data set of mangrove responses to N
fertilization experiments conducted for 3-12 years. N addition (900-2,700 kg N/tree/yr)
decreased survival probability by 10%, with trees growing on the land side in hypersaline
conditions particularly vulnerable when rainfall was low.
11.7.2.	Riparian Wetland
The Great Salt Lake is an inland salt lake ecosystem that provides an important food
source for migrating bird species as well as an economically important fishing industry
by providing habitat for Artemia franciscana, brine shrimp. Carling et al. (2013) studied
impounded and floodplain wetlands along the rivers that feed into the GSL, and found
that increasing nitrate and nitrite concentrations in water flowing through the wetlands
decreased reproduction of the submerged aquatic plants Ruppia cirrhosa and Stuckenia
spp. (equations in Table 11-8).
June 2018
11-73
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
11.7.3. Summary Table
Table 11-8 Nitrogen loading effects upon demography.
Type of
Ecosystem
Additions or
Load (kg N/ha/yr)
Biological and Chemical
Effects
Study
Site
Study
Species
Reference
Mangrove
No areal rate
given; 900 kg/yrto
2,700 kg N/yr per
tree as urea
Mangrove survival probability
decreased 10% with N addition.
12 global Mangrove
sites, species
including
Indian
River
Lagoon,
FL
Lovelock et al.
(2009)
Riparian
wetlands
and
floodplain
wetlands
Not reported
Reproduction (Repro, g/m2
drupelets) declines with
increasing surface water nitrate
(NO3 in mg/L):
Repro = -4.15 * ln(NC>3) + 4.03,
R2 = 0.51; as well as with
increasing surface water NO2"
(NO2" in mg/L):
Repro = -13.43 x ln(N02")
-25.27, R2 = 0.54).
Great
Salt
Lake, UT
Submerged
aquatic
vegetation:
Ruppia
cirrhosa and
Stuckenia
spp.
Carlinq et al. (2013)
g = gram; ha = hectare; kg = kilogram; L = liter; m = meter; mg = milligram; N = nitrogen; N02 = nitrite; N03 = nitrate; yr = year.
11.8. Biodiversity/Community
High biodiversity is a characteristic of wetlands. Fluctuating water levels in open systems
such as riparian strips and tidal marshes create zones with distinct abiotic characteristics
and multiple niches for microbes, plants, and animals with different biotic and abiotic
requirements. More closed systems, such as bogs and fens, are nutrient-poor,
high-organic-acid ecosystems that contain rare species including carnivorous plants,
ericaceous plants, and bryophytes characterized by low growth rates. Most wetlands
contain species with a range of N tolerance and N-acquisition strategies, and one result of
eutrophication is to shift the composition and relative abundance of species so that
nitrogen-tolerant species become more common as species that are adapted to low
nitrogen availability become less common or disappear from the system. Thus, plant
cover (percentage of the total wetland area occupied by a particular species) is an
important metric of biodiversity. Similarly, communities of aquatic producers contain
both nitrogen-tolerant and low-nitrogen-adapted species, and changes in phytoplankton
species composition represents a change in biodiversity important for dependent food
June 2018
11-74
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
webs (see Appendix 9.3.2. Appendix 10.3.2. and Appendix 10.3.3). Changes in the
abundance of herbivorous, omnivorous, or predatory invertebrates are indicative of
alterations to trophic interactions and effects, which could cascade up the food web
towards economically or culturally important fish, bird, and mammal species.
The 2008 ISA found that the evidence was sufficient to infer a causal relationship
between N deposition and the alteration of species richness, species composition, and
biodiversity in wetland ecosystems. The 2008 ISA noted there are 4,200 native plant
species in U.S. wetlands, 121 of which are federally endangered. Wetlands provide
habitat to a disproportionally high number of rare plants given their relative area; for
example, fens occupy 0.01% of the land area of northeast Iowa, yet contain 17% of the
endangered, threatened, or listed species of concern in the area. Wetland species have
evolved under N limited conditions, including endangered species in the Isoetes (three
endangered species) and Sphagnum (15 endangered species) genera, as well as the
endangered insectivorous plants Sarracenia oreophila and Drosera rotundifolia.
The 2008 ISA provided evidence from American bogs and fens that showed that
Sarracenia purpurea (purple pitcher plant) responded to N deposition with changes in
morphology and projected population extinction under N deposition levels of 4.5-6.8 kg
N/ha/yr. Coastal wetlands responded to N enrichment with increased primary production,
shifting microbial and plant communities, and altering pore water chemistry, although
many of the studies in coastal wetlands used N enrichment levels more similar to that of
wastewater than of atmospheric deposition.
11.8.1.
11.8.1.1.
Plants
Salt Marsh
Nitrogen was an important factor driving plant diversity in a study on the Patuxent River,
where marsh soil NOs -N and pore water salinity best explained the variation in plant
species richness out of all factors considered, with NO3 -N accounting for 26% of the
variation in the richness of all species. Presence-absence data for all species were used to
determine the geographic range for each species, and the 50% of species per river with
the smallest range was designated as a "restricted-range" subset. On the Patuxent River,
the best model of variation in species richness of restricted range species contained only
soil NO3 -N as an explanatory variable, with a positive relationship between NO3 -N and
species richness [see Table 11-9; (Sharpe and Baldwin. 2013)1.
June 2018
11-75
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Several new field addition studies further confirm N addition alters plant community
composition in salt marshes. In Kirkpatrick Marsh, MD, a limited plant community and
resampling of fixed plots allowed researchers to measure the biomass of plots as a proxy
for plant cover. C4 plants Spartina patens and Distichlis spicata plot cover increased over
the course of the 4-year fertilization experiment so that C4 biomass was 129% higher in
the 4th year than in the 1st year of receiving 250 kg N/ha/yr, while C4 biomass in the
control plots remained unchanged over 4 years. The C3 plant species Schoenoplectus
americanus declined in biomass in the control plots by 8-22% of that produced in the
1st year, but decreased to a greater extent in N addition plots, declining after 4 years by
52% compared to the 1st year plot biomass (Langlcv and Megonigal. 2010). In the
long-term nutrient addition experiment at Great Sippewissett Marsh, MA, 30 years of
addition of sewage sludge altered community composition. Nitrogen addition at 170, 520,
and 1,560 kg N/ha/yr altered the relative composition of Spartina alterniflora and
Distichlis spicata, with percentage cover of S. alterniflora decreasing 3% with every
additional 100 kg N/ha/yr, and percentage cover of D. spicata increasing 3% with every
additional 100 kg N/ha/yr of total N load (Fox et al.. 2012).
In Elkhorn Slough, CA, marsh community was assessed at six sites. At one site, where
impoundment restricted tidal exchange, N addition of 150 kg N/ha/yr altered community
composition, decreasing cover of native marsh plants by 52% and increasing cover of
non-native upland plants by 80% (Goldman Martone and Wasson. 2008). There were two
studies with N addition above 500 kg N/ha/yr (Baldwin. 2013; Ryan and Bover. 2012).
but this N addition level was not useful in inferring effects of N deposition upon
ecosystems.
11.8.1.2. Freshwater Tidal Marsh
In a freshwater tidal marsh on the Tchefuncte River, LA, N addition shifted the relative
dominance of the perennial wetland-obligate monocots dominant at the site (Graham and
Mendelssohn. 2010). Increasing N loads increased the relative dominance of Persicaria
punctata (formerly Polygonum punctatum) with a 0.13% increase (percentage of total
biomass consisting of P. punctata) for every 10 kg N/ha/yr added. Increasing N loads
decreased the relative dominance of Eleocharis fallax, with a 0.08% decrease (percentage
of total biomass consisting of E. fallax) for every 10 kg N/ha/yr added (Graham and
Mendelssohn. 2010) N loading did not affect the relative dominance of Sagittaria
lancifolia, but did alter its storage ofN in plant tissues (Appendix 11.5).
June 2018
11-76
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
11.8.1.3. Freshwater Wetland
There are two new studies of N loading in freshwater systems; they do not quantify N
deposition or N loading in terms of an areal rate, but rather find correlations between
surface water or soil N and changes in plant community diversity across a gradient of N
loads. Nitrogen loading is often correlated with invasions of non-native plants and
associated reductions in native species abundance and richness (Green and Galatow itsch.
2002). An observational study of 48 wetlands along the coasts of the Great Lakes found 7
floristically distinct plant communities: 1 Sphagmim-dominated wetland, 3 marsh types
dominated by native plants, and 3 marsh types dominated by invasive species. A
classification and regression tree that successfully classified 79% of the sites to the
correct plant community used surface water total N, conductivity, and pH as explanatory
variables: one of the invasive plant communities (dominated by the invasive Typha spp.)
was associated with high total N (Johnston and Brown. 2013).
In a similar observational study of 28 restored wetlands in Illinois, soil nitrogen
significantly correlated with plant invasion and a decline in native plant species richness
(Matthews et al.. 2009). There were linear, positive correlations of soil nitrate with
relative percentage cover of non-native plants, and with relative percentage cover of
invasive Phalaris arundinacea (see Table 11-9 for full equations). For each 0.1 ppm
increase in soil nitrate, non-native plant cover increased 4%, and P. arundinacea cover
increased 5%. The surveyed wetlands were diverse communities: there were 483 vascular
plant species across all 20 wetlands, and each wetland contained 40-100 native species.
For every 0.1 ppm increase in soil ammonium across the wetlands, seven native species
disappeared; and for every 0.1 ppm increase in soil nitrate, three native species
disappeared (Matthews et al.. 2009).
In addition to these studies, the National Wetlands Condition Assessment (NWCA),
conducted by U.S. EPA in 2011, collected soil and water chemistry data as well as plant
biodiversity data. In California wetlands sampled for this project, higher surface water
nitrate + nitrite concentrations correlated with lower wetland condition scores as
measured by the California Rapid Assessment Method (CRAM) (U.S. EPA. 2016i).
11.8.1.4. Intermittent Wetland
Vernal pools are intermittent wetlands, which typically flood in the late winter and
spring, and serve as important habitat for amphibians and endemic invertebrates.
California vernal pools host a number of endangered and threatened animal species
during their wet phase, as well as a high diversity of plant species during their dry phase.
June 2018
11-77
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
There are no studies published that directly test the effect of N deposition or N addition to
vernal pools, but there is a mesocosm study, which may allow some inference ofN
effects on vernal pools. This study added combined N and P solutions to mesocosms and
found increased algal biomass, which decreased vascular plant cover and richness in
vernal pool mesocosms (Kncitcl and Lessin. 2010). N and P additions were not quantified
in areal rates.
11.8.1.5. Bog and Fen
In Mer Bleue Bog, Ontario, N addition decreased the dominance of Sphagnum mosses in
the northern ombrotrophic peat bog. After 4 years of N fertilization at the rate of 64 kg
N/ha/yr, cover of Sphagnum species (of Sphagnum magellanicum and Sphagnum
capillifolium) in fertilized plots was 43% of Sphagnum cover in control plots (Juutinenet
al.. 2010). At Year 12, cover of Sphagnum species in 16 kg N/ha/yr fertilized plots was
64% of cover in control plots (Larmola et al.. 2013). Plots measured in the same growing
season (2011) that had received 7 years of elevated N at higher N addition rates also
experienced declines in Sphagnum cover: 26% decline in cover at 32 kg N/ha/yr and 54%
decline in cover at 64 kg N/ha/yr (Larmola et al.. 2013). The following year, moss
species (Sphagnum spp. and Polystrichum strictum) cover was 63% lower in 64 kg
N/ha/yr plots than in control plots, while the percentage of cover of the vascular plant
Chamaedaphne calyculata was 202% higher in 32 kg N/ha/yr plots than in control plots
(Wang et al.. 2016a). These declines in peat-forming species corresponded to changes in
ecosystem C fluxes (Appendix 11.3.2).
N deposition or addition effects upon plant community composition have been studied in
European systems. In an Irish fen, which was also the site of aN addition experiment (see
Appendix 11.4.4 and Appendix 11.5.5). researchers observed plant community
composition shifts in control plots over 4 years (Paulissen et al.. 2016). Under ambient
deposition of 7-10 kg N/ha/yr, cover of the N-sensitive brown moss Scorpidium
revolvens decreased 25%, while cover of peat moss Sphagnum contortum increased 20%
and cover of vascular plants increased qualitatively (Paulissen et al.. 2016). N addition
increases vacular plant cover in multiple systems. In an Italian fen, N addition altered
plant cover of 4 of the 17 species documented in the marsh (Gerdol and Brancaleoni.
2015). Vascular species cover did not shift in response to 10 kg N/ha/yr; but in response
to 30 kg N/ha/yr addition, cover of shrub Calluna vulgaris increased 96% and cover of
grass Molinia caerulea increased 5.6 times that of control plots. Like the vascular plants,
moss species Sphagnum fuscum (native in the U.S., listed as endangered by North
Carolina) did not respond to 10 kg N addition, but its cover decreased 45% in response to
30 kg N addition. The moss species Polytrichum strictum (native in the U.S., listed as
June 2018
11-78
DRAFT: Do Not Cite or Quote

-------
1	endangered in Kentucky) increased in response to both 10 or 30 kg N addition, with
2	cover 190-240% higher after 8 years of N addition. Researchers continued to monitor the
3	plots for 3 years after ceasing N addition treatments to assess wetland recovery: vascular
4	plants and P. strictum remained dominant, while S. fuscum cover continued to decline,
5	and a different moss species, Sphagnum magellanicum, increased its coverage in plots
6	recovering from elevated N addition (Gcrdol and Brancaleoni. 2015).
11.8.1.6. Summary Table
7
Table 11-9 Nitrogen loading effects upon plant biodiversity and communities.
Type of
Additions or
Biological and



Ecosystem
Load (kg N/ha/yr)
Chemical Effects
Study Site
Study Species
Reference
Coastal high
Addition: 150 kg
At a site where tidal
Six sites at
Marsh
Goldman
marsh
N/ha/yr as urea
inflow was restricted,
Elkhorn
community
Martone and

S dep = not
reported
N dep = not
reported
fertilization increased
Slough,
dominated by
Wasson

cover of non-native
Watsonville,
Sarcocornia
(2008)

upland plants by 80%
CA
pacifica, also


and decreased cover of

contains Jaumea


native marsh plants by
52%.

carnosa,
Frankenia salina,
Spergularia
salina, Distichlis
spicata, and
Atriplex
californicaAriang
uiaris

Estuarine salt
Addition: 1,337 kg
S. pacifica cover
China Camp
Sarcocornia
Rvan and
marsh
N/ha/yr as urea
increased at 6.2x the
State Park,
pacifica
Bover (2012)

S dep = not
rate of increase in
CA
dominant (C3


reported
N dep = not
reported
control. J. carnosa cover

succulent shrub),


declined 5.4x as fast as
in control.

Distichlis spicata
(C4 grass), and
Jaumea carnosa
(C3
semisucculent
forb)

Estuarine salt
Addition: 250 kg
S. americanus
Kirkpatrick
Schoenoplectus
Lanalev and
marsh
N/ha/yr
aboveground biomass
Marsh, MD
americanus (C3),
Meqoniqal

S dep = not
reported
N dep = not
reported
decreased by 19 and
(measured in
Spartina patens
(2012. 2010)

45% in the 3rd and 4th
3rd and 4th
(C4), and


yr, respectively.
yr,
Distichlis spicata



2008-2009)
(04)

June 2018
11-79
DRAFT: Do Not Cite or Quote

-------
Table 11-9 (Continued): Nitrogen loading effects upon plant biodiversity and
communities.
Type of
Ecosystem
Additions or
Load (kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species Reference
Coastal salt
marsh
Addition: 170 kg
N/ha/yr, 520 kg
N/ha/yr, 1,560 kg
N/ha/yr, all added
as sewage sludge
(10% N, 6% P2O5,
4% K2O)
S dep = not
reported
N dep = not
reported
Background total
N
load = 2 kg/ha/yr
(as reported in
Bowen and Valiela
(2001).
S. alterniflora cover (y, as
% cover) declines in plots
with increasing N load
(Nadd, as kg N/ha/yr),
y = -0.0292 x Nadd +
64.6.
D. spicata cover (z,as %
cover) increases in plots
with increasing N load,
z= 0.0327 x Nadd + 3.04.
Great
Sippewissett
Salt Marsh,
MA
Salt marsh
community,
dominated by
Spartina
alterniflora,
Spartina patens,
Distichlis spicata,
and Iva
frutescens.
Fox et al.
(2012)
Estuarine tidal
marsh
Addition: none
S dep = not
reported
N dep = not
reported
On the Patuxent River,
pore water NCV-N and
salinity best explained
plant species richness
(model R2 = 0.67), with
NC>3"-N accounting for
26% of variation in the
full data set and for 5% of
variation in widely
distributed species. In
species with more
restricted geographic
ranges, N03"-N alone
best predicted plant
species richness
(positive relationship,
model R2 = 0.71).
Nanticoke
River and
Patuxent
River,
Chesapeake
Bay, MD and
DE
Plant community
Sharpe and
Baldwin
(2013)
Estuarine tidal Addition: 670 kg
marsh	N/ha/yr
S dep = not
reported
N dep = not
reported
Typha spp. cover
increased by 120%.
Nanticoke
River, MD
and DE
Plant community Baldwin
(2013)
Freshwater
estuarine
marsh
50, 200, or
1,200 kg N/ha/yr
as Nutralene
methylene urea
S dep = not
reported
N dep = not
reported
N addition (Nadd as kg
N/ha/yr) decreased the
dominance of E. fallax
(Efa, as % biomass)
Elfa = -0.00812 x
Nadd + 15.410.
Addition increased the
dominance of P.
punctatum (Ppu, as % of
biomass)
Ppu = 0.03144 x Nadd + 9
.353
Tchefuncte
River,
Madisonville,
LA
Oligohaline plant
community
dominated by
Sagittaria
lancifolia,
Eleocharis fallax,
and Polygonum
punctatum
Graham and
Mendelssoh
n (2010)
June 2018
11-80
DRAFT: Do Not Cite or Quote

-------
Table 11-9 (Continued): Nitrogen loading effects upon plant biodiversity and
communities.
Type of
Ecosystem
Additions or
Load (kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species Reference
Restored
freshwater
wetlands
Ambient
deposition; rates
not reported
Soil available N (ppm
ammonium, abbreviated
as Nnh4, or ppm nitrate,
as Nno3) correlated with
decreasing native
species richness
(number of species, y):
y = -73.8A/a/h4 +122,
r2 = 0.41
y = -30.9Nno3 + 97.4,
r2 = 0.18
Soil available N (as ppm
nitrate, or Nno3),
correlated with increasing
cover by all nonnative
plants (relative % cover
all non-natives, z):
z = AMNno3 -3.5,
r2 = 0.31
Soil available N (as ppm
nitrate, or Nno3),
correlated with increasing
cover by invasive
Phalaris arundinacea
(relative % cover,
abbreviated PHAR):
PHAR= 52.7/Va/03-29.6,
r2 = 0.58
28 wetlands
restored or
created in
1992-2002,
Illinois
All wetlands
surveyed
summer
2006
Vascular plant
community:
483 species
across all
wetlands,
including
109 non-native
species
Matthews et
al. (2009)
Ombrotrophic
peat bog
16 (low), 32
(medium), or 64
(high) kg N/ha/yr
as N (NH4NO3) or
NPK (NH4NO3 and
KH2PO4)
S dep = not
reported
N dep = 8 kg
N/ha/yr
In high N treatment,
Sphagnum cover
declined by 57%.
In all NPK treatments,
Sphagnum cover
declined over 8 yrto
1-25% of control, as
Polytrichum cover
increased by 5-1 Ox
control and then
declined.
Mer Bleue
Bog,
Ontario,
Canada
Bog plant
community:
dwarf shrub
species and
mosses:
Sphagnum
magellanicum,
Sphagnum
capillifolium, and
Polytrichum
strictum
Juutinen et
al. (2010)
Ombrotrophic
16 (low), 32
After 12 yr of low N,
Mer Bleue
Shrub species
Larmola et
peat bog
(medium), or 64
Sphagnum cover
Bog,
(Vaccinium
al. (2013)

(high) kg N/ha/yr
decreased 36%. After
Ontario,
myrtilloides,


as N (NH4NO3) or
5 yr of medium and high
Canada
Ledum


NPK (NH4NO3 and
N, Sphagnum cover

groenlandicum,


KH2PO4)
decreased 26 and 54%,
respectively.

Chamaedaphne
calyculata) and
mosses
(Sphagnum
magellanicum,
Sphagnum
capillifolium,
Polytrichum
strictum)

June 2018
11-81
DRAFT: Do Not Cite or Quote

-------
Table 11-9 (Continued): Nitrogen loading effects upon plant biodiversity and
communities.
Type of
Ecosystem
Additions or
Load (kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species Reference
Ombrotrophic
peatland
Addition = 16 kg
N/ha/yr (5N, or 5x
background
deposition), 32 kg
N/ha/yr (1 ON), or
64 kg N/ha/yr
(20N), all as
NH4NO3
Ambient (wet)
deposition = 8 kg
N/ha/yr and up to
0.26 kg P/ha/yr
N addition increased
cover of Chamaedaphne
calyculata by
202%(1 ON), and
decreased cover of moss
species by 63% (20N).
Mer Bleue
bog, Ottawa,
Canada
Evergreen
shrubs:
Chameadaphne
calyculata and
Rhododendron
groenlandicum
dominant; sparse
Kalmia
angustifolia
Deciduous
shrubs: Vacciniu
m myrtilloides
Mosses:
Sphagnum
capillifolium, S.
magellanicum,
and Polystrichum
strictum
Wang et al.
(2016a)
Calcareous,
rich fen
Addition = 50 kg
N/m2/yr as NO3" or
50 kg N/m2/yr as
NH4+
Ambient
deposition =
7-10 kg N/m2/yr,
based on Aherne
and Farrell (2002)
Sphagnum cover
increased, Scorpidium
cover decreased, and
vascular plant cover
increased in all plots
(including control plots)
over the course of the
study (2000-2004).
Scragh Bog,
Central
Ireland
Scorpidium
revolvens and
Sphagnum
contort um
Paulissen et
al. (2016)
June 2018
11-82
DRAFT: Do Not Cite or Quote

-------
Table 11-9 (Continued): Nitrogen loading effects upon plant biodiversity and
communities.
Type of
Ecosystem
Additions or
Load (kg N/ha/yr)
Biological and
Chemical Effects
Study Site Study Species Reference
Fen
(transitional
mire)
Treatment,
2002-2009,
solution of
NH4NO3 added at
10 kg N/ha/yr
(Low N) or 30 kg
N/ha/yr (High N)
Deposition = not
reported
C. vulgaris cover does
not change in response
to low N addition, but 8 yr
of high N increased cover
(# intercepts) 96%, and
cover did not change 3 yr
after cessation of N
treatment (recovery
period).
S. fuscum cover does not
change in response to
low N, but 8 yr of high N
decreased cover (%)
45%, and cover
continued to decline
during recovery, when
cover was 60% less than
in control plots.
P. strictum cover
responded to 8 yr of low
and high N with
190-240% cover
increase, and although
this response declined in
recovery period, when
cover was 120-140%
higher than in control
plots.
M. caerulea cover did not
change in response to
low N, but increased to
5.6x control cover with
high N, and was 6.7x
control cover during
recovery period.
S. magellanicum cover
did not respond to low or
high N during 8 yr
experiment, but in
recovery, cover in plots
that had previously
received high N
increased to 4.3x control
plot cover.
Torbiera di
Passo San
Pellegrino,
Italy
Hummocks and
lawns:
13 species,
Calluna vulgaris,
Carex nigra,
Carex pauciflora,
Carex rostrata,
Eriophorum
vaginatum,
Molinia caerulea,
Polytrichum
strictum,
Sphagnum
angustifolium,
Sphagnum
fuscum,
Sphagnum
magellanicum,
Sphagnum
russowii,
Vaccinium
uliginosum,
Vaccinium vitis-
idaea
Lawns had an
additional
four species:
Nardus stricta,
Potentilla erecta,
Trichophorum
cespitosum,
Vaccinium
myrtillus
Gerdol and
Brancaleoni
(2015)
11.8.2. Consumers
Increased reactive N can alter dynamics among consumers in wetlands. Studies show that
increasing N load increases parasitism, decreases overall health condition, and increases
June 2018	11-83	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
the abundance of mosquitos (vectors for zoonotic diseases). There is also a new
meta-analysis evaluating consumer behavior in coastal wetlands. He and Silliman (2015)
conducted a meta-analysis of herbivory in globally distributed coastal wetlands (marshes
and mangroves) and found that N and N + P additions generally increased herbivore
abundance and herbivory across salt marsh studies, but not across mangrove studies.
In a survey of the native mud snail Ilyanassa obsoleta collected from 15 New England
salt marshes, increased N correlated with increasing parasitism of I. obsoleta. Sediment N
correlated positively with parasitic trematode prevalence (If = 0.41) and with infection
by the parasitic flatworm Stephanostomum spp. (If = 0.42).
Wigand et al. (2010) developed a metric for marsh condition using the Nitrogen Loading
Model (Wigand et al.. 2003) and data collected on abundance and richness of plants,
invertebrates, and soil conditions of 10 coastal marshes in Narragansett Bay, RI. Marsh
condition declined as annual N load to the marsh increased.
Ialeggio and Nvman (2014) collected plants dominant in freshwater (Panicum
hemitomon), freshwater-brackish (Sagittaria lancifolia), and brackish (Spartina patens)
marshes, fertilized the plants with 619 kg N/ha/yr in mesocosms, and then offered the
plants along with unfertilized controls in feeding trials to the introduced aquatic rodent
Myocastor coypus (nutria). Fertilization increased plant tissue N by 140%, from 0.9% N
to 2.2% N by mass, and nutria preferentially fed on fertilized plants. N loading increased
plant mass loss due to nutria herbivory by 750% (Ialeggio and Nvman. 2014).
N addition increases the abundance of mosquito species Culex tarsalis and Anopheles
hermsi, which act as vectors for zoonotic diseases. Duguma and Walton (2014)
constructed freshwater wetland mesocosms at UC Riverside, planting Schoenoplectus
maritimus in all mesocosms and monitoring plant growth as well as colonization of
mesocosms by mosquito species. N loads were 0, 32, 65, and 108 kg N/ha in 2009, and 0,
59, 119, and 198 kg N/ha in 2010. Sampling retrieved elevated numbers of Culex tarsalis
(western encephalitis mosquito) larvae and pupae in elevated nitrogen treatments in both
2009 and 2010, and of Anopheles hermsi in 2010. In 2009, immature mosquito counts
were 56-100% higher in fertilized treatments than in control treatments, but there were
no differences among different N loads in mosquito counts. In 2010, total mosquito
counts were similar in unenriched mesocosms and in mesocosms receiving 59 kg
N/ha/yr, but increased by 64% with 119 kg N/ha/yr, and by 129% with 198 kg N/ha/yr
(Duguma and Walton. 2014).
11-84
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
11.9.
Critical Loads
In the 2008 ISA, no critical loads studies had been published. Since then, critical loads
have been published for wetlands in the U.S. (Greaver et al.. 2011).
11.9.1. Freshwater Wetland
There are two published critical loads for North American freshwater wetlands. Greaver
et al. (2011) determined that the critical load for altered peat accumulation and NPP is
between 2.7 and 13 kg N/ha/yr, based on observations from Aldous (2002a). Moore et al.
(2004). Rochefort et al. (1990). and Vitt et al. (2003). The upper end of this critical load
range is based on measurements of wet deposition only [10 to 13 kg N/ha/yr; (Aldous.
2002a. b) | and, therefore, does not reflect total N loading from all sources. There is
evidence showing that N deposition alters both the morphology and population dynamics
of the purple pitcher plant. The empirical evidence suggests a critical load to protect the
population of purple pitchers of 10-14 kg N/ha/yr (Gotelli and Ellison. 2006). while
matrix modeling to forecast long-term population sustainability based on observations of
population demographics suggests a lower value of 6.8 kg N/ha/yr (Gotelli and Ellison.
2002).
11.9.2. Coastal Wetlands
Coastal wetlands have open hydrologic and nutrient cycles that receive N loads from
sources other than atmospheric deposition, so a critical load for atmospheric N deposition
has been difficult to establish (Greaver et al.. 2011). Typically, the amount of N added in
experimental treatments in these systems simulates total N input and, therefore, far
exceeds the atmospheric deposition received by U.S. coastal wetlands. Only two studies
have considered N loads below 100 kg N/ha/yr. Based on the results ofWigand etal.
(2003). a critical load to protect the community structure of salt marshes is likely to be 63
to 400 kg N/ha/yr. Caffrev et al. (2007) provided additional evidence that 80 N/ha/yr
alters microbial activity and biogeochemistry. Greaver et al. (2011) also established a
critical load for eelgrass growing in estuarine and marine waters associated with coastal
wetlands (see Appendix 10.6). The critical load for coastal wetlands is based on total N
loading to salt marshes. Total loading includes N deposition directly to the marsh surface,
N deposited to the watershed and transported via surface or groundwater, and runoff from
agriculture, urban areas, or other sources. Additional experimental evidence on
ecosystem response to N loads more similar to N loading from atmospheric deposition is
needed to improve the critical load calculation for coastal wetlands in the U.S.
June 2018
11-85
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
11.9.3.
Comparison to Critical Loads from Europe
An assessment of N critical loads in Europe was published in 2003 and designated critical
loads for multiple types of wetlands, including raised and blanket bogs, poor fens, rich
fens, mountain rich fens, and intertidal wetlands (Bobbink et al.. 2003). There are
numerous publications on N effects to wetlands in Europe compared to the U.S. In
general, documented responses include effects on growth and species composition,
competition among species, peat and peat water chemistry, decomposition, and nutrient
cycling. A brief summary of the European critical loads for wetlands is presented here.
Bobbink et al. (2003) assigned a critical load of 5 to 10 kg N/ha/yr for bog ecosystems,
based on plant community and species responses to N deposition, and indicated that
precipitation and P limitation should be used to assign critical loads to individual sites.
The observed changes in the plant communities of ombrotrophic bogs included the
replacement of Sphagnum-form i ng species with nitrophilous moss species [20 to
40 kg N/ha/yr in Dutch bogs; (Grcvcn. 1992)1 and the absence of characteristic
Sphagnum species in British bogs [30 kg N/ha/yr; (Lee and Studholme. 1992)1. A recent
study along a deposition gradient in Alberta, Canada, compared Sphagnum bog and fen
responses in that region to previously published European bog and fen responses to N and
S deposition. Similar N deposition gradients elicited significant Sphagnum and water
chemistry responses in European peatlands but no responses in Albertan peatlands
(Wieder et al.. 2016). This suggests that critical loads for European Sphagnum species,
which have experienced a longer history of higher N deposition than North American
wetlands, may not be applicable in North American bogs and fens. Bobbink et al. (2003)
also suggested a critical load for the carnivorous bog plant roundleaf sundew
[10 kg N/ha/yr in Swedish bogs; (Bobbink et al.. 2003; Redbo-Torstensson. 1994)1 that
may be relevant to American bogs where this species also grows. The European critical
load for this carnivorous plant is similar to the range of critical loads suggested for
freshwater wetlands in the U.S.
There are European critical loads for fens, but they are based on moss decline, and the
amount of N causing ecological effects may not be directly relevant to U.S. ecosystems
for the reasons outlined by Wieder etal. (2016). European poor fens have a critical load
of 10 to 20 kg N/ha/yr based on increased sedge and vascular plants and negative effects
on mosses. The critical load for rich fens in Europe is 15 to 35 kg N/ha/yr based on
increased tall graminoids and decreased diversity. The critical load for montane rich fens
in Europe was 15 to 25 kg N/ha/yr based on increased vascular plants and decreased
bryophytes. These community changes are similar to the increasing vascular plant and
declining moss cover observed in American N addition experiments (see
Appendix 11.10). although there has been no critical load established for American bogs
June 2018
11-86
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
and fens based on these community composition changes. In Europe, changes in the
vegetation composition and structure likely affect fauna species assemblages, such as
ground-breeding birds, spiders, and beetles, living in the originally open bog vegetation.
Increased nutrient availability results in an increase of the nutrient content of plant
material (Limpens et al.. 2003a; Tomassen et al.. 2003) and algal growth (Limpcns et al..
2003b; Gulati and Demott. 1997). which affects herbivorous, detritivorous, and
carnivorous invertebrates (van Duinen et al.. 2004).
The European critical load for salt marshes, based on expert judgment, is 30 to
40 kg N/ha/yr (Bobbink et al.. 2003). but studies of European salt marshes are limited.
There are no European wetland systems composed of native plants directly equivalent to
North American low salt marshes; in European systems, these zones consist of
unvegetated mud flats, or of monocultures of invasive plants. High levels of N input (65
to 70 kg N/ha/yr) significantly increased biomass production in the Netherlands (van
Wijnen and Bakker. 1999); however, no changes in species composition and in diversity
have been observed at deposition of 15 to 25 kg N/ha/yr at sites in the Netherlands and
Germany (Bobbink et al.. 2003). The critical load for N deposition in North American
coastal wetlands may be close to these values when considered as part of the total N load
that these systems receive.
11.10. Summary
There are chemical and biological effects of N deposition in wetlands. Wetland
vegetative communities are adapted to high levels of natural organic acidity, so it is
unlikely that S or N deposition would cause any acidification-related effects at levels of
acidic deposition commonly found in the U.S. IYU.S. EPA. 2008a). in Annex B],
Wetlands can be sensitive to eutrophication caused by N deposition. These effects were
documented in the 2008 ISA. In this appendix, new information published since 2008 is
integrated with the information published in the 2008 ISA. Synthesis is made across
wetland types for causality statements. Information is also synthesized for freshwater and
coastal wetlands because saltwater and estuarine wetlands are typically adapted to much
higher N loading.
11.10.1. Causality across Wetland Types
New studies published between 2008-2015 support and extend the findings of the 2008
ISA. As in the 2008 ISA, the body of evidence is sufficient to infer a causal
relationship between N deposition and the alteration of biogeochemical cycling in
June 2018
11-87
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
wetlands. N deposition changes N and C cycling in wetlands. There is new evidence of
N deposition changes to plant physiology, which was not included in the 2008 ISA as an
endpoint due to lack of evidence. This evidence expands our understanding of the causal
relationship between N deposition and species diversity. The body of evidence is
sufficient to infer a causal relationship between N deposition and the alteration of
growth and productivity, species physiology, species richness, community
composition, and biodiversity in wetlands.
N deposition contributes to total N loading in wetlands, with the relative contribution of
deposition to total loading variable among wetland types. Chemical indicators of N
deposition in wetlands include NOs and NH4+ leaching, DON leaching, N
mineralization, denitrification rates, and N2O emissions. A wetland can serve as a source,
sink, or transformer of atmospherically deposited N (Dcvito et al.. 1989). Source/sink N
dynamics in wetlands vary with season, hydrological conditions, vegetation type, climate,
and surface geology (Mitchell et al.. 1996; Arhcimcr and Wittgrcn. 1994; Koerselman et
al.. 1993; Devito et al.. 1989).
Several new studies provide evidence of N deposition alterations to biogeochemical
cycling of N within a wetland ecosystem (Appendix 11.3). A synthesis study of multiple
wetland types across the globe shows wetland N removal is proportional to N load (log[N
removal] = 0.943 x log[N load] - 0.033 N) and removal efficiency is 26% higher in
nontidal than tidal wetlands (Jordan et al.. 2011). Meta-analysis of 19 N addition
observations (N addition 15.4 to 300 kg N/ha/yr) found that N enrichment increased
wetland N2O emissions by 207% (Liu and Greaver. 2009). The emission factor for the
amount of N added to a wetland that is converted to N2O ranges between 0.007-0.07 (Liu
and Greaver. 2009).There are also new studies that evaluate the effects of N loading/N
addition on other endpoints related to N cycling in peat bog, riparian, mangrove, and salt
marsh wetlands. The endpoints evaluated include ecosystem N retention, wetland export
of N to surface waters, N fixation, N mineralization, denitrification, emission ofN20; and
bacterial abundance, activity, and composition in wetland soils. The results of North
American studies are summarized in Figure 11-2. Across studies, N loading decreases the
ability of wetlands to retain and store nitrogen, which may diminish the wetland
ecosystem service of improving water quality.
11-88
DRAFT: Do Not Cite or Quote

-------


¦f N03- and NH4+ in peat A
M
O
CQ
sl/ N retention efficiency B
0
c
.ro
fU
Ł sl/ BNF rates, sl/N-fixing symbionts c
100

o
bO
™ 'T* Denitrification rate D
100
"5 n)/ N mineralization E
fU
\1/ N retention tidal N export) F
-t—'
1/1
fU
° \1/ N retention (^porewater NH4+, ^N20 emissions) G
P. australis expansion and^porewater NH4+ h
100
180
250
250
0 100 200 300 400 500
N addition (kg N/ha/yr)
BNF =biological nitrogen fixation; ha = hectare; kg = kilogram; N = nitrogen.
aPinsonneault et al. (2016).
"Xing et al. (2011).
cRuess et al. (2013).
dWhiqham et al. (2009).
eVivanco et al. (2015).
fBrin et al. (2010).
gPastore et al. (2016).
"Mozdzer et al. (2016).
Figure 11-2 Summary of the levels of nitrogen addition where a change to
nitrogen cycling is observed.
1	There is new information on how N deposition alters biogeochemical cycling of C in
2	wetlands. In the 2008 ISA, evidence from Canadian and European peatlands showed that
3	N deposition had negative effects on Sphagnum bulk density and mixed effects on
4	Sphagnum productivity depending on the history of deposition. In Canadian
5	ombrotrophic peatlands experiencing deposition of 2.7-8.1 kg N/ha/yr, peat
6	accumulation increased with N deposition, but accumulation rates had slowed by 2004,
7	indicating a degree of N saturation (Moore et al.. 2004). N addition alters belowground
8	and aboveground pools of carbon and also increases wetland methane emissions, as
June 2018
11-89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
summarized by meta-analysis in the 2008 ISA and published in Liu and Greaver (2009).
In a study of 90 wetlands around the Gulf Coast from Florida to Texas, Nestlerode et al.
(2014) found that higher N in the soil correlated with lower bulk density (ln[soil
%N] = -1.9233 x [g/cc] + 0.4165) which indicates that the marsh is less resilient to
physical stresses from tidal or storm flooding, and the marsh platform is more likely to
shear off or wash away. New evidence shows that N loading increases methane
production from wetland soils. In N addition experiments replicated in three California
marshes (Vivanco et al.. 2015; Irvine et al.. 2012). field-measured methane flux from
soils increased linearly with increasing N load (0, 100, 200, 400, 800, 1,600,
3,200 kg N/ha/yr), such that CH4 flux from soils increased by 1.23 |ig CHVmVday for
each 10 kg N/ha/yr added, (Vivanco et al.. 2015; Irvine et al.. 2012). Methane flux
increased from low or nearly negative to net positive values above -100 kg N/ha/yr,
indicating a shift from net methanotrophy to methanogenesis in the microbial community
in response to N loading consistent with other studies (Liu and Greaver. 2009).
Additional literature evaluates the effects of N deposition, N loading, or experimental N
addition on C cycling in salt marshes, mangroves, freshwater tidal marshes, riparian or
intermittent marshes, bogs, and fens. Significant effects of N loading upon
biogeochemical cycling of C in North American wetlands (in which the N addition was
500 kg N/ha/yr or lower) are summarized in Figure 11-3. Ecological endpoints evaluated
to assess N loading effects on C cycling include (1) measures of C pools—plant
aboveground biomass and productivity, plant belowground biomass of roots and
rhizomes, soil organic matter, and total soil C or peat C; (2) measures of C fluxes—
microbial mineralization, decomposition rates, CO2 emissions, and CH4 emissions; and
(3) measures of physical marsh stability as a proxy for long-term C storage—peat
chemical composition, temperature, bulk density, physical resistance, and soil water
content. In general across wetlands, nitrogen loading increases aboveground C pools,
decreases or does not change belowground C pools, and does not change or increases C
fluxes. This shift from belowground to aboveground C storage may diminish the wetland
ecosystem services of long-term carbon storage and flood protection, as well as reduce
the stability and persistence of wetlands on the landscape. N loading effects upon C
cycling are uneven across species, which may affect wetland biodiversity and the wetland
ecosystem service of provisioning.
11-90
DRAFT: Do Not Cite or Quote

-------
s
Ł*5
*
AG: 4-3-21% growth 5. purpurea (STE)
BG: 4" 17% biomass in top 10 cm peat
BG: P and K limitation of microbial activity
BG: altered peat temperature, •!¦¦ peat [C02]
AG: ^ 30% biomass of moss community
AG: 1s 105% biamass C. catyculata (STE)
AG: -T1 600% biomass C. oligosperma (STE)
AG: i" 2.6 times productivity plant community
BG: 4- 46% net ecosystem C exchange
BG: 4-51% established root biomass
ANPP responsiveto N
EG: 4^ 71% rhizome biomass
BG: 4^ 33% macro-organic matter
AG: ^ 1.4-3.8 times biomass Z. miliacea (STE)
A


B
0

C
0

C
0

c
0

a
0

0
0

E
0
E

60
I

60
F

[m
G


H


i
I.J


500
500 I
500
ip
5
Ł
Ł
AG: ^ biomass A. germinans
AG: ^ biomass of 5. depressa
M
AG: ^"3-14% shoot mass in high marsh
AG: 1s 36-53% biomass of 5. parifka*
BG: 4- 42-34% fine root production
O.P
i
BG: T* roots of invasive P. australis
p
AG: ^ 25-7596 biomass colonizing plants
AG: 'V129% biomass of C4 plants S. patens and D. spicataC
AG: 4^55% bioma5sofC3 5. americanusC
AG: ^ 47-79% plant community biomass*
AS
[250J
f==°1
S
I 250 I
N addition (kg N/ha/yr)
AG = aboveground; ANPP = aboveground net primary productivity; BG = belowground; C02 = carbon dioxide; FW = freshwater;
ha = hectare; kg = kilogram; N = nitrogen; NPP = net primary productivity; STE = state-listed threatened or endangered species;
yr = year.
ACrumlev et al. (2016); BWendel et al. (2011); cPinsonneault et al. (2016); DXing et al. (2011); Elversen et al. (2010); FLarmola et al.
(2013); GGraham and Mendelssohn (2016); "Graham and Mendelssohn (2010); 'Ket et al. (2011); JFrost et al. (2009); KWhiqham et
al. (2009); LVivanco et al. (2015); "(Johnson et al.. 2016a); "Goldman Martone and Wasson (2008); "Lanalev and Meaoniaal (2010);
pLanalev and Meaoniaal (2012); "Mozdzer et al. (2016) RLanalev et al. (2013); sLanalev et al. (2009).
Grey boxes with black borders represent ambient deposition studies, and white boxes with blue borders represent field nitrogen
addition experiments.
Figure 11-3 Summary of new literature of nitrogen load effects on
belowground and aboveground carbon cycling.
June 2018
11-91
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
N addition effects on plant stoichiometry and plant architecture were not addressed in the
2008 ISA. Plant stoichiometry theory considers the balance of multiple chemical
elements in living tissues, and plant architecture is defined as the three-dimensional
organization of the plant body. The stoichiometry of plant tissue is often connected to the
tissue's physiological function. Most new studies on physiology and stoichiometry have
been conducted in bogs and fens, where N addition typically caused an increase in plant
N content, a decrease in N use efficiency and resorption, and an increase in plant
production, particularly of vascular plants. At high N loads or cumulative exposure to
years of lower N loads plants may experience leafN saturation and micronutrient
limitation (e.g., P, K, and Ca, indicated by altered leaf tissue concentrations or altered
reabsorption efficiencies; see Figure 11-4). which in turn cause leaf damage (Bubicr et
al.. 2011; Xing etal.. 2011) or decreasing plant abundance of sensitive species (see
biodiversity section). N loading can disrupt nutrient acquisition of carnivorous plants via
their adaptation of capturing and digesting insects; for example, along a narrow N
deposition gradient (3.4-5.0 kg N/ha/yr) across bogs in the Adirondack Mountains,
purple pitcher plant (S. purpurea) experienced negative effects upon growth and
insectivory at deposition of 4.4-4.9 kg N/ha/yr (Crumley et al.. 2016). A new European
study also found negative effects of N deposition upon insectivory of the wetland plant
Drosera rotundifolia at 3.81-11.30 kg N/ha/yr (Millctt et al.. 2012). Historically low N
loads to bogs and fens have made these wetlands and their endemic plant species
particularly sensitive to N deposition. New studies of physiology and plant architecture in
North American riparian wetlands, freshwater tidal marshes, mangroves, and salt marshes
showed that N addition in these systems increased plant tissue N with a cascading effect
that increased plant primary production and changed plant architecture. N loading
increased plant height in salt and freshwater marshes, which Deegan et al. (2012) showed
made salt marshes top-heavy and resulted in a loss of marsh platform stability, with and
shredding and loss of marsh surface. In general effects of N loading upon physiology and
architecture vary by species across North American wetlands. Even species tolerant of N
loading may experience negative effects if N loading causes physiological limitation by
other nutrients, including Ca, P, and K. Plant physiology and architecture altered by N
loading may affect wetland biodiversity and resiliency to other disturbances.
11-92
DRAFT: Do Not Cite or Quote

-------
4/40% insectivory in S. purpurea (STE) |	|
^ 27% leaf N in shrub C. calyculata (STE) | |
•4/ 15% leaf C:N, 4/ 34% leaf Ca, ^ amino acids in shrub C. calyculata (STE) [»D
4' 23% leaf Ca, 4' 55% P, 4^ 47% K in shrub V. myrtilloides (STE) 1161
^ 100% Vcmax in shrub R. groenlandicum (STE) QU
4^29-62% N uptake in moss S. magellanicum 1201
4/ 34% tissue Ca in mosses 1321
1* 42% P resorption in shrub C. calyculata (STE) \El
"Is 12 times P resorption, ^ 92-123% K resorption in R. groenlandicum (STE)
4/ 81% NUE, 4^ 91% NRE in shrub C. calyculata (STE)	[io]
4^ 89% NUE, 4/ 84% NRE in sedge C. oligosperma (STE)	| 601
¦T N uptake by grass C. canadensis	1601
4/45% NRE in tree A. rugosa	| 601
4^ 42%tissue Ca, "T 39%tissue P in mosses	1641
^ 16% leaf N, 4^ 54% leaf P, /t" glutamic acid, in shrub R. groenlandicum (STE)	1641
4/ 23% leaf Mg, 4- 186% Mg resorption in R. groenlandicum (STE)	1641
^ 84% leaf chlorophyll, /\* amino acids in shrub C. calyculata (STE)	1641
¦T P resorption 33%, 4/ 22% leaf Ca, 4/ 45% Mn in shrub C. calyculata (STE)	1641
Shift in bog seasonal gross PSN: 4^ summer PSN, ^ fall PSN	1641
10% leaf N , 4/ 21% P resorption in tree A. tenuifolia
AG plant tissue [N], in sedge B. maritimus (STE) 1321
planttissue N, plant N:P, 4^ NRE, 4^PRE in forb S. lancifolia	1501
"T 26% BG plant tissue [N] in sedge B. maritimus (STE)	1651
5% leaf construction cost in native haplotype of grass P. australis
25-43% leaf N, 4^ 18-28% leaf C:N, 1s 25-61% leaf N:P in tidal FW marsh
^ leaf N in succulent forb S. depressa	| ioo|
0	100
kg N/ha/yr
AG = aboveground; BG = belowground; C = carbon; Ca = calcium; FW = freshwater; K = potassium; Mg = magnesium;
Mn = manganese; N = nitrogen; NRE = nitrogen resorption efficiency; NUE = nutrient use efficiency; P = phosphorus; PP = primary
productivity; PRE = phosphorus resorption efficiency; PSN = photosynthesis; STE = state-listed threatened or endangered species;
Vcmax = maximum velocity of carboxylation.
Grey boxes with black borders represent ambient deposition studies, and white boxes with blue borders represent field nitrogen
addition experiments.
Figure 11-4 Summary of the level of nitrogen load that caused a change in the
response variables of plant stoichiometry and physiology in
wetlands.
1
2	The 2008 ISA found that the evidence was sufficient to infer a causal relationship
3	between N deposition and the alteration of species richness, species composition, and
4	biodiversity in wetland ecosystems. Wetlands provide habitat to a disproportionally high
5	number of rare plants given their relative area; for example, fens occupy 0.01% of the
6	land area of northeast Iowa yet contain 17% of the endangered, threatened, or listed
7	species of concern in the area. The 2008 ISA noted that there were 4,200 native plant
June 2018
11-93
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
species in U.S. wetlands, 121 of which are federally endangered. Wetland species have
evolved under N limited conditions, including endangered species in the Isoetes
(3 endangered species) and Sphagnum (15 endangered species) genera, as well as the
endangered insectivorous plants Sarracenia oreophila and Drosera rotundifolia. Plant
demography was not specifically addressed in the 2008 ISA, although demography is an
important aspect in maintaining species richness and biodiversity. Plant demography is
the change in plant population size and structure through time. In bogs and fens, N
addition experiments suggest positive population growth rates for Sarracenia purpurea at
0 or 1.4 kg N/ha/yr, but population losses at 14 kg N/ha/yr (Gotclli and Ellison. 2006).
Demographic modelling of S. purpurea found that populations remained stable for
100 years at deposition rates of 4.5-6.8 kg N/ha/yr, but extinction risk for populations
increased above 6.8 kg N/ha/yr (Gotclli and Ellison. 2002). Field N addition studies
suggest that N addition to bogs and fens will also affect community composition
(Figure 11-5). In Mer Bleue Bog, Ontario, N addition decreased the dominance of
Sphagnum mosses in the northern ombrotrophic peat bog, while increasing dominance of
woody shrub species (Larmola et al.. 2013; Juutinen et al.. 2010). This result is consistent
with the earlier European literature summarized in the 2008 ISA, which showed that N
deposition decreased moss dominance and increased the dominance of vascular plants in
European bogs and fens.
Across freshwater wetlands, N load is correlated with an increase in the abundance of
invasive plant species and a decrease in the number of native plant species (see
Figure 11-1 and Appendix 11.8.1). as well as an increase in larvae of mosquito species
that are vectors for zoonotic diseases (Appendix 11.8.2). In a freshwater tidal marsh on
the Tchefuncte River, LA, N addition shifted the relative dominance of the perennial
wetland-obligate monocots dominant at the site (Graham and Mendelssohn. 2010). In salt
marshes, N addition caused negative reproductive effects in invasive Spartina foliosa and
Spartina hybrids (Tyler et al.. 2007). and a positive effect on reproduction in Salicornia
bigelovii (Bover and Zedler. 1999). N loading in salt marshes is correlated with species
richness. A study on the Patuxent River found marsh soil NOs -N and pore water salinity
best explained variation in plant species richness out of all factors considered (R2 = 0.67),
with NO3 -N accounting for 26% of the variation in total species richness., In the same
system, the best model for rare species richness contained only soil NO3 -N as an
explanatory variable, with a positive relationship between NO3 -N and species richness
[i?2 = 0.71 (Sharpe and Baldwin. 2013)1. In salt marshes in Maryland and California, N
addition was shown to alter the relative abundance of plant species (Langlev and
Megonigal. 2010; Goldman Martone and Wasson. 2008) (Figure 11-5). N deposition
effects can cascade beyond plants up trophic levels to consumers. Meta-analysis showed
that N loading increased invertebrate abundance and herbivory, and a field experiment
found increased herbivory by the invasive mammal nutria in freshwater tidal and salt
11-94
DRAFT: Do Not Cite or Quote

-------
1	marshes (Appendix 11.8.2). In New England, N loading decreased salt marsh condition,
2	evaluated by a multidimensional metric that included abundance and richness of plants
3	and invertebrates as well as soil conditions (Appendix 11.8.2). In general, N loading
4	across North American wetlands decreases richness or abundance of sensitive species
5	while promoting invasive species, with negative effects on biodiversity.
CuO
o
CO
tc
o
u
4/ 36% cover of mosses S. magelj^inicum and...
1* 202% cover of shrub C. calyculata (8 yr Ex) B
4/ 26% cover of mosses S. mageHanicum and...
4/ 63% cover of mosses (8 yr Ex) B
4/ 54% cover of mosses S. mage^lanicum and...
4/ 57% cover of mosses S. magellarjicum and S....
4/ relative dominance of sedge E. fallax D
'T relative dominance of forb P. punctata D
/t" 80% cover of non-native upland plant species E
4/ 52% cover of native marsh plant species E
'T 129% cover of C4 grasses S. patens and D....
4/ 52% cover of C3 sedge S. americanus F
100
250
250
200 300
kg N/ha/yr
400
500
Ex = experimental exposure length; FW = freshwater; h = hectare; kg = kilogram; N = nitrogen; yr = year.
ALarmola et al. (2013).
BWang et al. (2016a).
cJuutinen et al. (2010).
"Graham and Mendelssohn (2010).
EGoldman Martone and Wasson (2008).
FLanalev and Meaoniaal (2010).
Numbers indicate the lowest addition level in which change is observed.
Figure 11-5 Summary of nitrogen addition studies on wetland biodiversity.
June 2018
11-95
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
11.10.2. Coastal versus Freshwater Wetlands
Coastal and freshwater wetlands tend to have different sensitivity to N
deposition/loading. The effect of N deposition on wetland ecosystems depends on the
fraction of rainfall in the total water budget, and the sensitivity to N deposition has been
suggested as: bogs (70-100% rainfall) > fens (55-83% rainfall) > coastal wetlands
[10-20% rainfall (Morris. 1991)1.
Greaver et al. (2011) suggested a critical load to protect biodiversity and biogeochemistry
of coastal wetlands based on salt marsh community composition, microbial activity, and
biogeochemistry (63-400 kg N/ha/yr as total N load). A comparison of those critical
loads with data on N addition levels (15-500 kg N/ha/yr) and associated effects,
published since the last ISA, is given in Figure 11-6.
In freshwater systems, Greaver et al. (2011) determined a critical load for wetland C
cycling as quantified by altered peat accumulation and NPP, between 2.7 and
13 kg N/ha/yr. Greaver et al. (2011) also set a critical load to protect biodiversity based
on morphology and population dynamics of the purple pitcher plant (Sarracenia
purpurea) between 6.8-14 kg N/ha/yr. A more recent study across an N deposition
gradient suggests that purple pitcher plant populations experience negative effects of N
deposition at 4.4 kg N/ha/yr (Crumley et al.. 2016). A comparison of freshwater wetland
CLs to data from N addition levels (16-500 kg N/ha/yr) and associated effects, published
since the CLs for wetlands were set, is given in Figure 11-7. There is information on the
relationship between N addition and numerous endpoints. At the lowest experimental
addition level (16 kg N/ha/yr), there are observations of altered C and N cycling and
altered biodiversity. The endpoints affected include decreases in moss cover, increased
peat biomass, decreased N retention efficiency, and altered/damaged leaf stoichiometry in
vascular plants.
June 2018
11-96
DRAFT: Do Not Cite or Quote

-------

aj



o
Denitrification rate
\wo]

c
ra
2
AG: ^ biomass A. germinans
1100 I


4- N mineralization
[ 100 I


't methane emissions
1100 I


AG: biomass of succulent forb S. depressa
1ml


Is leaf N in S. depressa
1100 I


1s shoot mass in high marsh
140


AG:biomass of S. pacifica
150


'T" cover of non-native upland plant species
150

-C
4- cover of native marsh plant species
150

ra
5
4- N retention ("T1 tidal IM export)
M

a:
ra
4- N retention (^porewater NH4+, *tN20 emissions)
250

l_)
BG: 4, native fine root production
jlso'


BG: T" roots of invasive P. australis
no


AG: 1s plant community biomass
250


AG: T1 biomass of C4 plants S. patens and D. spicata
I2501


AG: 4- biomass of C3 S. americanus
(13


colonization by invasive P australis
HED


"t1 cover of C4 grasses S. patens and D. spicata
250


4/ cover of C3 sedge S. americanus
0§]
a: 3
-a 2


t- Ł
ID "C
U
3
c
Community structure, microbial activity, biogeochemistry
63-400
0 100 200 300 400 500
kg N/ha/yr
AG = aboveground; BG = belowground; CH4 = methane; GTR = general technical report; N = nitrogen.
Figure 11-6 Summary of field nitrogen addition studies for coastal wetlands (blue
borders) versus critical load (black border).
June 2018
11-97
DRAFT: Do Not Cite or Quote

-------
BO .52
O J3 ¦
CO g
a 2'
4^growth of S. purpurea
4*40% insectivory in S. purpurea (STE)
1s 27% leaf N in shrub C. calyculata (STE)
4, N retention efficiency of bog
peat DIN
BG: biomass in top 10 cm peat
BG: recalcitrantC in peat
BG: T microbial metabolism of C
BG: P and K limitation of microbial activity
BG: altered peat temperature, 4- peat [C02]
AG: 1* biomass of moss community
4-leaf C:NP 4-leaf Ca, "t1 amino acids in shrub C. calyculata...
4, leaf Ca, 4/leaf P, 4'leaf K in shrub V. myrtilloides (State T& E)
1"" Vcmax in shrub R, groenlandicum (STE)
4-cover of mosses S. magellanicum and capillifolium(12 yr Ex)
4' N uptake in moss S. magellanicum
4>tissue Ca in mosses
1" P resorption in shrub C. calyculata (STE)
resorption, 'T* K resorption in R. groenlandicum (STE)
4' cover of mosses S. magellanicum and capillifolium (7 yr Ex)
4* NUE, 4- NRE in shrub C. calyculata (STE)
4' NUE, 4- NRE in sedge C. oligosperma (STE)
1" N uptake by grass C. canadensis
4- NRE in tree A, rugosa
AG: 't biomass of grass C. calyculata (STE)
AG: T biomass of sedge C. oligosperma (STE)
AG: "T productivity plant community
4/tissue Ca, 'f tissue P in mosses
^ leaf N, 4-P, 'T glutamic acid, in R. groenlandicum (STE)
4' leaf Mg, 4. Mg resorption in R. groenlandicum (STE)
T* leaf chlorophyll, amino acids in shrub C. calyculata (STE)
1" P resorption, 4- leaf Ca, 4' Mn in shrub C. calyculata (STE)
BG: 4» net ecosystem C exchange
Shift in bog seasonal gross PSN: 4' summer PSN, 1" fall PSN
4' cover of mosses S. magellanicum and capillifolium (4 yr Ex)
4' BNF rates, 4-N-fixingsymbionts
leaf N, 4- P resorption in tree A. tenuifolia
4/ SLM in tree A. tenuifolia
I" AG plant tissue [N], in sedge B. maritimus (STE)
"T larval and pupal mosquito abundance
T plant tissue N, T N:P, 4' NRE, 4-PRE in forb S. lancifolia
4- relative dominance of sedge E. fallax
'f relative dominance of forb P. punctata
BG plant tissue [N] in sedge B. maritimus (STE)
"T total mosquito abundance
ANPP responsive to N
^ leaf construction cost of grass P. australis
BG: 4- rhizome biomass
BG: 4' macro-organic matter
AG: -f biomass Z. miliacea(STE)
T1 leaf N, sb leaf C:N, 1" leaf N:P in tidal FW marsh
peat accumulation and NPP
Pitcher plant community change
16
16
16
16
16
16
16
16
16
16
16
16
|20|
I<200I
1250]
kg ^/fia/yr
AG = abovegrourid; ANPP = aboveground net primary productivity; BG = beiowgrourid; BNF = biological nitrogen fixation;
C = carbon; Ca = calcium; C02 = carbon dioxide; Ex = experimental exposure length; FW = freshwater; GTR = general technical
report; K = potassium; Mg = magnesium; Mn = manganese; N = nitrogen; NPP = net primary productivity; NRE = nitrogen resorption
efficiency; NUE = nutrient use efficiency; P = phosphorus; PRE = phosphorus resorption efficiency; PSN = photosynthesis;
SLM = specific leaf mass; STE = state-listed threatened or endangered species; Vcmax = maximum carboxylation velocity; yr = year.
Values indicate biotic or chemical changes observed in response to experimental nitrogen addition (boxes with blue borders) or an
ambient gradient of N deposition (boxes with black borders).
Figure 11-7 Summary of nitrogen load studies for freshwater wetlands as well
as current critical loads.
June 2018
11-98
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
APPENDIX 12. NONACID SULFUR ENRICHMENT
EFFECTS
Appendix 12 summarizes research on the nonacidifying impacts of S deposition on
ecosystems, synthesizing literature published since the 2008 ISA for Oxides of Nitrogen
and Sulfur-Ecological Criteria (hereafter referred to as the 2008 ISA) with earlier key
studies. The causal statements are presented in the Introduction (Appendix 12.1).
Appendix 12.2 discusses the effects of sulfur (S) deposition on S storage and cycling in
ecosystems, sulfide phytotoxicity in wetlands, internal eutrophication in freshwater
aquatic systems, methane emissions from wetlands and lakes, and the microbial
communities responsible for methanogenesis.
Increasing bioavailability of mercury (Hg) due to S addition is a major focus, and Hg
sources, pools, and fluxes in terrestrial and aquatic ecosystems are briefly described in
Appendix 12.3.1 and in the Supplemental Material. Appendix 12.3.2 includes a
discussion of the current understanding about the distribution of Hg methylation potential
across the prokaryotic domains and the key role of sulfur-reducing prokaryotes (SRPs).
Mercury methylation occurs in lakes in shallow sediments or below the oxycline; in peat
wetlands, at the interface with upland areas and generally near the water surface in peat;
in periphyton; in saturated decomposing litter or algal biomass; in saturated forest soils;
and in estuarine and marine sediments. In freshwater systems, methylation is strongly
seasonal, with methylmercury (MeHg) concentrations peaking in summer or fall.
Methylation rates and MeHg concentrations are determined by the biological niche and
environmental requirements of microbial methylators; the current state of our
understanding of those environmental requirements is briefly summarized in
Appendix 12.3.3.
In Appendix 12.3.4. multiple lines of evidence are presented establishing the relationship
between SOx deposition and increases in bioavailable Hg in the environment, including
experimental S addition experiments and case studies from well-studied wetlands where
S inputs are primarily agricultural. Appendix 12.3.5 considers observational evidence of
the correlations between SOx deposition and Hg burdens in fish. Appendix 12.3.5 also
presents the results of observational studies in prairie pothole wetlands, peat bogs,
freshwater marshes, streams, and rivers that show correlations between ambient sulfate
and MeHg concentrations in water and sediment samples. Appendix 12.4 summarizes
studies that find higher Hg burdens in mosquitos and fish from ecosystems with increased
experimental or anthropogenic S loading. Ecosystems especially sensitive to the effects
of S deposition are described in Appendix 12.5.
July 2018
12-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Appendix 12.6 describes critical loads for Hg from European systems; at this time, there
are no established critical loads for nonacid S effects in North American ecosystems.
Appendix 12.7 is a summary including causal determinations based on the synthesis of
new information and previous evidence.
Introduction
In the 2008 ISA, the evidence was sufficient to infer a causal relationship between S
deposition and increased methylation of Hg in aquatic environments where the value of
other factors was within adequate range for methylation. Direct atmospheric deposition of
S to wetland and aquatic systems, as well as sulfate leaching from terrestrial systems with
current or historical S deposition, alters biology and chemistry of these systems.
Nonacidifying effects of S deposition in these nonterrestrial systems include toxicity to
wetland plants, increased phosphorus (P) availability leading to internal eutrophication,
altered methane emissions, and increased Hg methylation and Hg concentration in biota.
Consistent with the causal statement in the 2008 ISA, the body of evidence is sufficient
to infer a causal relationship between S deposition and the alteration of Hg
methylation in surface water, sediment, and soils in wetland and freshwater
ecosystems. The 2008 ISA described sulfide phytotoxicity in European systems, but
there is more recent research demonstrating sulfide phytotoxicity under current
conditions in North American wetlands. The body of evidence is sufficient to infer a
causal relationship between S deposition and changes in biota due to sulfide
phytotoxicity including alteration of growth and productivity, species physiology,
species richness, community composition, and biodiversity in wetland and
freshwater ecosystems.
S deposition contributes to Hg accumulation by stimulating the activity of
anoxic-sediment-dwelling sulfur-reducing bacteria (SRB), which convert inorganic Hg to
MeHg in the course of metabolism. The 2008 ISA described the activity of SRB, but
more recent research indicates that S reducing archaea are also active in wetland
sediments, which is why this document will use the broader term sulfur-reducing
prokaryotes (SRPs) to denote both bacteria and archaea involved in S reduction.
Laboratory and mesocosm-scale experiments reviewed in the 2008 ISA established that
only trace amounts of MeHg could be produced in the absence of sulfate. These results
were confirmed with larger scale observational studies. MeHg enters the food chain and
bioaccumulates in higher trophic levels, and research summarized in the 2008 ISA
suggested that numerous wild populations of fish, birds, and mammals experienced
MeHg exposures high enough to cause substantial reproductive, behavioral, or health
impairment.
12-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
At the time of the 2008 ISA, dose-response relationships between S deposition and
Hg-methylation rates had not been established, in part because oxygen content,
temperature, pH, and labile carbon supply also control the rates of the activity of SRB in
aquatic environments. Watersheds with conditions known to be conducive to Hg
methylation were identified in the northeastern U.S. and southeastern Canada, although
significant biotic Hg accumulation had been observed in other regions that had not been
studied as extensively. The U.S. EPA set the fish tissue criterion in 2001 for MeHg at
300 ng/g fish tissue (reported as 0.3 mg/kg) for the protection of human health, which
resulted in 2,436 fish consumption advisories for Hg in 2004, 2,682 in 2005, and 3,080 in
2006. Forty-eight states, one territory, and two tribes had issued Hg fish-consumption
advisories at the time of the 2008 ISA.
Appendix 12 summarizes research on the nonacidifying impacts of S deposition on
ecosystems, synthesizing literature published since the 2008 ISA with earlier, key studies
(Figure 12-1).
CH4 = methane; Fe = iron; Hg = mercury; MeHg = methylmercury; P = phosphorus; S042 = sulfate; SOx = sulfur oxides.
Figure 12-1 Effects of sulfur oxide deposition on chemical (blue boxes),
biological (green boxes), and atmospheric (yellow boxes)
indicators of ecosystem change, as documented by the previous
Integrated Science Assessment and more recent research.
July 2018
12-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
12.2.
Ecosystem Effects of Altered Sulfur Cycling
This section provides background on the sulfur cycle in terrestrial and aquatic ecosystems
(Appendix 12.2.1). It describes effects of S enrichment (SOx deposition or experimental
S amendments) upon aquatic S cycling (Appendix 12.2.2). sulfide toxicity on wetland
plants (Appendix 12.2.3). S enrichment on phosphate cycling in aquatic ecosystems and
on uptake of toxic elements by aquatic plants (Appendix 12.2.4). and S enrichment in
altering microbial competition and methanogenesis (Appendix 12.2.5).
12.2.1. The Sulfur Cycle
S cycling in terrestrial ecosystems is relatively well characterized compared to S cycling
in aquatic systems. Sulfate is the dominant S source in supplying plant nutrient S
demand. Sulfate in the soil solution, whether its source is weathering of sulfur-containing
minerals, mineralization of internal ecosystem S, or atmospheric deposition of SOx, has
several possible routes through the S cycle (Figure 12-2).
Wet
Deposition
Dry
Deposition
Litter
Inputs
Uptake
Desorption
Source: Adapted from Mitchell et al. (2011).
Figure 12-2 Sulfur cycle in terrestrial, forested ecosystems.
July 2018
12-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
Sulfate can be immobilized by incorporation into living cells (microbes, fungi, plants),
which ultimately end up in the organic soil S pool. In most North American terrestrial
ecosystems, total S pools in soil greatly exceed S stored in living biota (Figure 12-2).
with soil S in organic or inorganic molecules. The largest pool of S in soil is organic,
typically comprising more than 95% of total soil S in temperate ecosystem soils (Wang et
al.. 2006; Solomon et al.. 2001). although in drained wetland soils of the Florida
Everglades Agricultural Area (EAA), the organic fraction was quantified as 87% of total
S (Ye et al.. 2010). Organic S pools in grassland soils are negatively correlated with mean
annual temperature in the North American Great Plains, indicating that climate is a strong
control on S cycling (Wang et al.. 2006). Organic soil S is found in ester-bound forms
(produced via microbial activity) or bound directly to carbon (C) in organic compounds
of varying sizes [whose ultimate sources were found to be plant roots or litter, (Edwards.
1998)1. Organic soil S cannot be taken up by plants directly, but organic S compounds
are mineralized back to sulfate by soil microbes under favorable moisture, temperature,
and oxic conditions (Wang et al.. 2006). Organic S pools can be indirectly enhanced by
SOx deposition via microbial and plant incorporation, as indicated by measures of soil
[S] and S isotopes. Sulfate in the soil solution can remain in the soil S pool in its
inorganic form as it adsorbs to hydrated soil particles. Soil adsorption of sulfate generally
results in a smaller S pool than soil organic matter, but this pool can be directly enhanced
by SOx deposition and associated acidification. Acidification may increase the storage
capacity of soils for sulfate. In the case of nonspecific adsorption, increasing acidity and
more H+ held in colloidal suspension create more sulfate binding capacity in soil, while
ecosystem recovery towards a more neutral pH may cause desorption of sulfate and
higher sulfate concentrations in the soil solution. In the case of specific adsorption,
sulfate binds with metal atoms in mineral particles; sulfate specific adsorption to soil also
increases with lower pH. However, the rate at which sulfate bound to soil particles is
released from adsorption in the soil solution as the ecosystem recovers (see Appendix 4)
towards a more neutral pH will heavily depend upon soil factors such as clay mineralogy,
soil age and weathering, calcium availability, content of amorphous iron and aluminum
oxides, and glaciation history. Inorganic S storage within terrestrial soils can be directly
affected by SOx deposition, as indicated by measures of sulfate concentrations in soil
fractions and soil solution.
Sulfate in the soil solution may be transported out of the terrestrial ecosystem if it leaches
to surface water and into downstream wetland and aquatic ecosystems. Sulfate
concentrations in surface water are often measured as an indicator of S fluxes from small
terrestrial watersheds. In watersheds that have received historically high S deposition,
dissolved organic sulfur (DOS) may also comprise a substantial portion of the S-leaching
from terrestrial to aquatic watersheds. In Archer Creek in the Adirondacks, NY, DOS
averaged 21% of total dissolved sulfur in the creek (Kang et al.. 2014). Freshwater
12-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
aquatic S cycling is not as well described in the scientific literature as is terrestrial soil S
cycling. The preponderance of evidence is concerned with the locations of redox S
reactions within these aquatic systems. Sulfate leached from terrestrial ecosystems, as
well as sulfate from direct deposition to surface water, is transformed in aquatic systems
under particular conditions. High oxygen levels in the water column or inundated
sediments of a wetland, lake, or stream favor S in oxidized form as sulfate. Sulfate in
anoxic (very low O2) waters or sediments is transformed into more reduced forms such as
sulfide or elemental S, which are less soluble and may precipitate into sediments.
Sulfur-reducing prokaryotes (SRPs) couple their metabolism to the transformation of
sulfate into these reduced forms. SRPs are obligate anaerobes but require sulfate for
metabolism, and as a result, they are often present and active at oxic-anoxic boundaries in
aquatic environments. In shallow lakes, this boundary (and the majority of sulfate
reduction) occurs at the water-sediment interface, in the top 1-2 cm of sediment (Rudd et
al.. 1986; Kelly and Rudd. 1984). In deeper lakes that experience seasonal stratification,
the water column may account for up to 15% of sulfate reduction (Ingvorscn and Brock.
1982). and the anoxic boundary may occur in the flocculant layer (suspended sediment);
for example, the boundary occurs 10 cm above the top of the sediment in Little Rock
Lake, WI (Sherman et al.. 1994). In wetlands, particularly in peat bogs that depend
entirely upon precipitation as their water supply, the anoxic-oxic boundary often
coincides with the water table and can be much more dynamic. Finally, recent research
suggests that periphyton occupying high-oxygen waters can create anoxic
microenvironments that promote the growth and activity of SRPs (see Appendix 12.3.3).
The location and activity of SRPs are important environmentally because some SRPs
couple sulfate reduction with Hg methylation (see Appendix 12.3.2). The transformation
of S between reduced and oxidized forms in aquatic and wetland ecosystems occurs in
particular locations within these ecosystems but can have broader impacts on chemistry
and biota.
Pools of S in freshwater systems (lakes, rivers, streams, and wetlands) and residence
times of S in ecosystem pools are addressed in only a few ecological studies. S cycling in
lakes removes sulfate from the water column and stores reduced S in sediments. Sulfate
in the water column diffuses into sediments where it binds to particulate matter or is
incorporated by microbes. Water column sulfate is reduced (by chemical or biological
processes) and released into the sediments as sulfide, which can bind to iron and
precipitate, or as elemental S, which can precipitate directly. An early study by Rudd et
al. (1986) developed a budget for reduction-oxidation cycling of S based on Adirondack
lakes, and estimated that 47% of water column sulfate was reduced and stored in
sediments. A broader study that included lakes in the Adirondacks, Ontario, the
Experimental Lakes Area, Northern Wisconsin, and Southern Norway estimated that
39-80% of annual sulfur load was retained in sediments based on mass balance, or
12-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
11-65% based on measured sulfate reduction rates, across all lakes (Kelly et al.. 1987).
On average across lakes, one-third to one-half of annual lake sulfate load was transferred
to sediments, and variation around this mean correlated with water residence times of
lakes. Longer hydraulic residence times correlated with more sulfate removal. The
release of reduced S into the water column is responsible for some of the other
nonacidifying effects of S deposition, including internal eutrophication and protective
effects against certain heavy metals (see Appendix 12.2.4). The rest of the S originally
taken up by microbes as sulfate is incorporated into organic compounds. The growth and
activity of SRPs are responsible for Hg methylation in many sediments (see
Appendix 12.3). About one-third of the reduced inorganic S and 57% of the organic S are
buried in lake sediments annually, or about 47% of the total sulfate input. The remaining
reduced sulfur, 53% of the original sulfate input, is oxidized by chemical reaction with
water molecules (Bates et al.. 2002) and returns to the S cycle (Rudd et al.. 1986). These
studies suggest that changes in S inputs to lakes—for example, reductions in SOx
deposition—could rapidly reduce sulfate concentrations in the water column, assuming
other environmental factors remain steady (see below for climatic influences on S
cycling).
Wetland S budgets resemble lake budgets in S reduction and burial of reduced S in
sediments, but also may include plant uptake and S export if hydrology allows. In a study
of depressional freshwater wetlands in two lakes in Germany, the sulfate reduction rate
was 2.5 to 7.9 times higher annually in the high sulfate lake (22.7 mg sulfate/L) than in
the low sulfate lake (9.0 mg/L) (Kleeberg et al.. 2016). Sulfur budgets have also been
determined for the peat soils and freshwater marshes of the Florida Everglades. The EAA
has large impacts on S cycling in freshwater marshes in the surrounding designated
Water Conservation Areas (WCA), which serve as a water quality buffer around
Everglades National Park, a Class I area. In the EAA, the addition of elemental S as an
agricultural soil amendment increased sulfate in the soil from 13 to 19% of total soil S,
although at the end of the growing season sulfate was 1% of total S due to uptake by
growing sugarcane (Ye et al.. 2010) and runoff into surrounding canals and wetlands
(Bates et al.. 2002). In an early study quantifying S budgets in the Everglades WCA,
(Bates et al.. 2002) determined that most of the sulfate in the system has its source in S
agricultural amendments applied in the EAA. In these wetland systems, the pool of S in
organic matter is high, and additional sulfate could be rapidly incorporated into organic
matter or exported in surface water, in addition to the reduction or burial, which occurs in
other aquatic systems.
Weather and climatic factors can disrupt S storage in more stable pools within aquatic
and wetland systems. In smaller stratified lakes, sulfate transfer from the epilimnion into
lower layers occurs slowly by diffusion between the stable water layers, while in larger
12-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
lakes, the layers are more prone to perturbation by wind and allow larger transfers of
sulfate to zones where SRPs are active (Ingvorsen and Brock. 1982). Sulfur cycling, like
most microbially driven processes, is seasonal; sulfate fluxes to the sediment are 10 times
higher in the summer than in winter (Sherman et al.. 1994). and the growing season
coincides with peak sulfate reduction in many freshwater systems (see
Appendix 12.3.3.3V In the Experimental Lakes Area (ELA) in Ontario, 10-39% of
reduced S ended the summer as sulfide dissolved in the hypolimnion, and this S was
reoxidized to sulfate when circulation occurred in the fall (Kelly et al.. 1982). In winter,
O2 can penetrate deeper into sediments or peat, reoxidizing the sulfide (Sherman et al..
1994). This reoxidation can also occur in water bodies where water levels are controlled
by humans, for example in reservoirs (Ecklev et al.. 2015). Extreme events like droughts
can also result in reoxidation of reduced S (Wasik et al.. 2015) and can contribute to Hg
methylation hotspots upon rewetting and restoration of anaerobic aquatic conditions.
Deposition and Sulfur Stores
In terrestrial ecosystems, atmospheric deposition of S at levels below amounts that trigger
acidification may still induce biochemical effects. These effects were described in the
2008 ISA and include nutrient enrichment of plants, increased S storage in soils, and
elevated sulfate concentrations in water leaching from soils. New information regarding
the effects of S deposition on terrestrial biogeochemical cycling is sparse, but can be
found in Appendix 4. Terrestrial plant productivity is commonly limited by N, so SOx
deposition at levels low enough to provide S enrichment without associated acidification
effects is unlikely to have significant effects upon terrestrial plant productivity. Instead,
the nonacidifying effects of SOx deposition are strongest in freshwater systems such as
wetlands and lakes, where hydrology controls storage and release of S.
Changes in water availability and water levels in aquatic systems alter the S cycle and
may release additional S to the aquatic environment from storage in sediment. During a
drought at Little Rock Lake in Wisconsin, lake sulfate concentrations increased
0.18 mg/L/yr, from 1.5 mg/L in 1998 to 2.9 mg/L in 2006 (Watras and Morrison. 2008).
During this time period, lake water volume decreased only 30%, which would not
account for the 93% increase in water sulfate concentration, suggesting that drought
released additional S stored in lake sediment or biomass into the water column. Watras
and Morrison (2008) calculated that an additional load of 5 kg S/ha/yr from internal
ecosystem sources of S would account for the increase of S in Little Rock Lake. In the S
addition experiment at the S6 bog in Marcell Experimental Forest, MN (see
Appendix 12.3.4.1). rewetting of the bog after a drought released a pulse of sulfate into
surface water in both control and experimental S addition plots. In plots to which
12-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
32 kg S/ha/yr had been added, concentrations of sulfate in peat pore water after the
drought were 438% higher than in control plots ("Wasik et al.. 2012). indicating that S
addition decreases the retention of S in drought-stressed wetlands (Table 12-1). This is
particularly important because intensification of drought frequency and severity is one of
the possible effects of climate change, and suggests that climate change and SOx
deposition may synergistically contribute to increased sulfate concentrations in surface
water and associated water quality impairment.
Table 12-1 New study on sulfur (S) deposition effects on sulfur (S) cycling.
Type of
Ecosystem
Additions or Load
(kg S/ha/yr)
Biological and
Chemical Effects
Study Site Study Species Reference
Bog	Ambient S deposition:
5.5 kg/ha/yr (mid 2000s) as
quantified at NADP site
MN16
S addition: 32 kg S/ha/yr
dissolved in pond water and
delivered by sprinklers
(mimics 4x the 1990s
deposition rate)
Following 2007
drought, pore water
SO42" concentrations
were 438% higher in
experimental
treatment.
S6 peatland,
bog section,
Marcell
Experimental
Forest, MN
Wasik et al.
Pore water,
peat, and Culex (2012)
spp.(mosquito)
larvae
ha = hectare; kg = kilogram; NADP = National Atmospheric Deposition Program; S = sulfur; S042 = sulfate; yr = year.
12.2.3. Sulfide Toxicity
Wetland ecosystems are characterized by occasional or total soil inundation and typically
include water and sediment zones with low oxygen levels. Under low-oxygen conditions,
sulfate can be used as an electron acceptor by a number of microorganisms and is quickly
converted to sulfide during decomposition of organic C. Sulfide inhibits plant root
nutrient uptake at high levels [>34.1 mg/L or >1 mM; (Koch etal.. 1990)1. and sulfide
toxicity due to increased sulfate loads was documented by the 2008 ISA and references
therein for wetland plants, including Carex spp., Juncus acutiflorus (not native to or
documented in North America), Galium palustre (endangered in Ohio, special concern in
Tennessee), other species within Gramineae (the grass family), as well as the aquatic
macrophytes Stratiotes aloides (listed as an Class C noxious weed in Alabama, Class 1
prohibited aquatic plant in Florida) and Elodea nuttallii (threatened in Kentucky, special
concern in Tennessee). As reviewed in the 2008 ISA, Smolders et al. (2003) set a
threshold value of <48 mg SO42 /L in surface water to protect the sensitive aquatic
July 2018
12-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
species Stratiotes aloides and Potamogeton acutifolius (not native to CONUS), as well as
to protect Potamogeton zosteriformis and Utricularia vulgaris, which are both native and
widely distributed in CONUS. This threshold would protect against sulfide phytotoxicity
and internal eutrophication and was based on correlations between surface water
concentrations and plant presence/absence in Dutch fens (Smolders et al.. 2003). A recent
survey of water quality and plant presence/absence data from four water management
districts in the Netherlands confirmed that the probability of encountering the S-sensitive
plants S. aloides and six S-sensitive Potamogeton (pondweed) species declined
significantly at sulfate concentrations greater than 50 mg/L (Vcrmaat et al.. 2016).
More recent research confirms sulfide toxicity in wetland habitats and suggests that
sulfide toxicity can determine plant community composition in freshwater wetlands. A
review of the sulfide toxicity literature by Lamers et al. (2013) found sulfide toxicity
occurred between 0.3-29.5 mg S27L (originally reported as 10-920 |imol L ') in
freshwater wetland emergent plants and aquatic submerged macrophytes native to North
America (see Table 12-2 for species-specific phytotoxicity levels). A recent study
sampled pore water chemistry and plant community composition in Junius Pond Fen,
Seneca County, NY, and found that sulfide concentrations (range: not detectable to
5.73 mg/L or 168 |iM H2S) had negative effects on the plant community. Sulfide
concentration was negatively correlated with total plant cover and had strong effects on
the cover of the dominant plants in the fen (Simkin et al.. 2013). In models, high sulfide
concentrations decreased cover of the moss species Campylium stellatum, decreased
cover of the monocot and USDA-classified (USDA. 2015b) wetland obligate Eleocharis
rostellata (endangered in Florida and Pennsylvania; threatened in Illinois, Minnesota, and
Wisconsin; sensitive or special concern in Rhode Island and Washington), and decreased
cover of obligate monocot Cladium mariscoides (endangered in Pennsylvania). Sulfide
concentrations also correlated negatively with dicot diversity (Simkin et al.. 2013). This
study shows that sulfide toxicity effects on sensitive plant species can cascade to affect
plant community composition in fens. There are also sensitive species in other freshwater
wetlands. A greenhouse experiment with Taxodium distichum (baldcypress) seedlings
collected from a North Carolina riparian field showed negative effects of sulfate addition
(sulfate concentrations: 48 mg/L in drought treatment, 129 mg/L in flooded treatment) on
seedling height, which the researchers attributed to sulfide toxicity (Powell et al.. 2016).
Baldcypress is a foundation species in southeastern freshwater swamps, which if located
in estuaries, may also experience elevated sulfate, sulfide, and salinity as a result of
saltwater intrusion. SOx deposition may exacerbate sulfide toxicity in baldcypress
swamps already stressed by sea level rise.
12-10
DRAFT: Do Not Cite or Quote

-------
Table 12-2 Quantitative effects of sulfide on wetland and aquatic plant species.
Species
Effects of Sulfide
Sulfide
(mg S27L)
Listing
Reference
Calla palustris
Decreased aboveground
productivity
4.8
NRCS wetland
obligate
Geurts et al., 2009 in
Lamers et al. (2013)
Caltha palustris
Decreased aboveground
productivity
5.5
NRCS wetland
obligate
Van der Welle et al.,
2007b in Lamers et
al. (2013)
Carex nigra
Decreased aboveground
productivity
0.3-0.6
NRCS wetland
facultative
Lamers et al., 1998 in
Lamers et al. (2013)
Ceratophyllum
demersum
Decreased aboveground
productivity
16
NRCS wetland
obligate
Geurts et al., 2009 in
Lamers et al. (2013)
Cladium jamaicense
Decreased leaf elongation
rate
Decreased net
photosynthetic rate
Aboveground die-off and root
(and rhizome) die-off
7.0
22.1
29.5
NRCS wetland
obligate
Li et al, 2009 in
Lamers et al. (2013)
Elodea nuttallii
Decreased aboveground
productivity
3.2
NRCS wetland
obligate
Van der Welle et al.,
2007a in Lamers et
al. (2013)
Eiodea nuttallii
Decreased aboveground
productivity
4.8-16
NRCS wetland
obligate
Geurts et al., 2009 in
Lamers et al. (2013)
Equisetum fluviatile
Decreased aboveground
productivity
1.6-16
NRCS wetland
obligate
Geurts et al., 2009 in
Lamers et al. (2013)
Juncus
alpinoarticulatus
seedlings
Decreased aboveground
productivity
1.0-1.6
NRCS wetland
obligate
Grootjans et al., 1997
in Lamers et al.
(2013)
Juncus effusus
Decreased aboveground
productivity
16
NRCS wetland
obligate or facultative
Geurts et al., 2009 in
Lamers et al. (2013)
Menyanthes trifoliata
Decreased aboveground
productivity
(unfertilized/fertilized)
4.8/>4.8
NRCS wetland
obligate
Geurts et al., 2009 in
(Lamers et al., 2013)
Menyanthes trifoliata
Decreased aboveground
productivity
>7.5
NRCS wetland
obligate
Armstrong and
Boatman, 1967 in
Lamers et al. (2013)
Nitella flexilis
Decreased aboveground
productivity
1.6

Van der Welle et al.,
2006 in (Lamers et
al., 2013)
July 2018
12-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
Table 12-2 (Continued): Quantitative effects of sulfide on wetland and aquatic
plant species.
Species
Effects of Sulfide
Sulfide
(mg S2"/L)
Listing
Reference
Panicum hemitomon
Decreased aboveground
productivity, root (and
rhizome) die-off
20.2
NRCS wetland
obligate
Koch and
Mendelssohn, 1989
in Lamers et al.
(2013)
Panicum hemitomon
Decreased aboveground
productivity, decreased root
ADH activity, decreased
nutrient uptake
32.1
NRCS wetland
obligate
Koch et al., 1990 in
Lamers et al. (2013)
Potamogeton
compressus
Decreased aboveground
productivity
4.8-16
NRCS wetland
obligate
Geurts et al., 2009 in
Lamers et al. (2013)
Sphagnum
cuspidatum
Aboveground die-off
1.9

Lamers et al., 1999 in
Lamers et al. (2013)
Thelypteris palustris
Decreased aboveground
productivity
4.8
NRCS wetland
obligate
Geurts et al., 2009 in
Lamers et al. (2013)
Typha domingensis
Decreased leaf elongation
rate, decreased net
photosynthetic rate,
aboveground die-off, root
(and rhizome) die-off
29.5
NRCS wetland
obligate
Li et al., 2009 in
Lamers et al. (2013)
Source: all columns except "Listing" are from Lamers et al. (2013).
Sulfide toxicity has been proposed as an important causal factor in the expansion of
Typha domingensis (southern cattail) in the Florida Everglades. Southern cattail is
displacing the historically dominant sawgrass, Cladium jamaicense, which is the
keystone species in the Everglades sawgrass prairies and which has minor agricultural
importance as a food source for small mammals and water birds (reported in. Everitt et
al.. 1999). In a hydroponic greenhouse study of both species, cattails had a higher
tolerance for sulfide in all plant metrics measured. Sulfide concentrations in growth
media were 0, 7.5, 15.7, 23.5, 31.4 mg/L sulfide (reported as 0, 0.22, 0.46, 0.69, or
0.92 mM). At 7.5 mg/L sulfide, sawgrass leaf elongation rates decreased 49%; at
15.7 mg/L sulfide, net photosynthetic rate decreased 34%; and at 23.5 mg/L, sawgrass
biomass in different compartments decreased 29-41% (Li et al.. 2009). Cattails
experienced sulfide toxicity effects on leaf elongation rates at 23.5 mg/L sulfide (38%
decrease) and on photosynthesis at 31.4 mg/L sulfide (22% decrease), with no effects of
sulfide upon biomass at any concentration (Li et al.. 2009). Together, these results
suggest that sulfide toxicity is shifting the competitive balance between these species
away from the dominant species in the Everglades. However, studies that assessed cattail
and sawgrass growth in mesocosms placed across the naturally occurring sulfate-sulfide
July 2018
12-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
gradient in the Everglades (DeBusk et al.. 2015) or in mesocosms with added sulfate (Li
et al.. 2009) have not found effects of sulfide toxicity upon sawgrass. Under controlled
conditions, sulfide toxicity favors cattails over keystone Everglades species sawgrass, but
evidence from field studies in the Everglades ecosystem suggests that geological or
ecological factors not reproduced in the greenhouse may mitigate sulfide toxicity.
Recent work by the Minnesota Pollution Control Agency (MNPCA) proposes to
incorporate geological and biological factors into water quality regulation in order to
prevent sulfide toxicity to an economically important species. Wild rice (Zizania
palustris) has been protected by the state of Minnesota as ecologically important food for
waterfowl and an economically important foraged crop (MPCA. 2015a). and the Ojibwe,
Menominee, and Dakota peoples value the plant as a cultural and economic resource
(Pastor etal.. 2017). The state first set a standard to protect wild rice in 1973, when the
water quality standard was set at 10 mg sulfate/L. In 2011, the state began collecting field
samples, conducting experiments in hydroponics and mesocosms, and modeling
environmental and experimental data in order to revise the standard. The previous
standard of 10 mg sulfate/L would only protect 58% of current wild rice sites from
sulfide toxicity (MPCA. 2015a). In March 2015, the Minnesota Pollution Control Agency
(MPCA) released its recommendations (MPCA. 2015a). Based on the new analyses, the
MPCA set a sulfide concentration threshold of 0.165 mg sulfide/L in sediment pore water
to protect wild rice, based on 2011-2013 monitoring of 112 Minnesota water bodies in
which this level of sulfide corresponded to a 10% decrease in the probability of the
species being present (MPCA. 2015b). Sulfide is the end product of microbial sulfate
reduction, which is heavily dependent upon other environmental factors, particularly
DOC, and sulfide phytotoxicity depends upon its residence time as a free ion in water,
which depends upon iron concentrations (see Figure 12-3). As a result, the MPCA did not
set a single value of sulfate as a standard to protect wild rice.
12-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
High iron
in sedime
High organic
carbon levels in
sediment
*	bacteria
| iron	iroiT)( iron^)
Iron in sediment binds to sulfide
and neutralizes it, making it
nontoxic to wild rioe.
Oganic carbon in sediment is
food forthe bacteria, causing
more sulfide to be produced.
Source: from MPCA (2015b).
Figure 12-3 Schematic from Minnesota Pollution Control Agency that
illustrates the mitigating effect of iron on the toxicity of sulfide
and the stimulatory effect that organic carbon has on sulfide
production.
Instead, MPCA established a formula that would allow protective sulfate thresholds to be
calculated for all wild rice-producing water bodies based on sediment organic carbon and
sediment iron concentrations in each water body [Source: (MPCA. 2015b)l:
Sulfate = 0.0000136 x Organic Carbon—1.410 x Ironl.956
Equation 12-1
The range in sulfate values that protect wild rice from sulfide toxicity in Minnesota water
bodies is from 0.4 to >200 mg SO-r /L. This proposed standard will require large-scale
sediment sampling to determine sediment pore water DOC and Fe concentrations in
Minnesota water bodies, and is still under review. These regulatory efforts determine
protective surface water sulfate concentrations on the basis of a sulfide phytotoxicity
threshold protective of wild rice. In hydroponic experiments with Z. palustris, sulfate and
sulfide exposures did not affect seed germination, but sulfide concentrations of 0.32 mg/L
or higher reduced growth rates 88% compared with controls (Pastor et al.. 2017). The
same research group also conducted growth trials in outdoor tank mesocosms in which
surface water sulfate concentrations, and thus sediment sulfide concentrations, could be
manipulated. In the mesocosms, 50 mg S0.f7L did not affect seedlings, but elevated
July 2018
12-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
sulfate and sulfide at 100 mg SO42 /L or higher decreased seed mass, seed viability,
seedling emergence rates, and seedling survival rates (Pastor et al.. 2017).
Sulfide phytotoxicity as a result of SOx deposition alters freshwater wetlands and lakes,
but is unlikely to affect coastal wetlands. Seawater contains high concentrations of sulfate
[1 ppt seawater contains 750 |iM or 72 mg/L sulfate according to (Hackney and Avery.
2015)1. and as a recent study in the Cape Fear Estuary, North Carolina, demonstrates,
seawater intrusion and sulfate-reducing conditions structure the distribution of tidal
marsh and tidal swamp zones in the estuary. Specifically, tidal marsh plants are salt and
sulfide tolerant, and occur in the estuary in areas where sulfate reduction dominates
microbial mineralization (Hackney and Avery. 2015). Tidal swamp plants (such as
baldcypress, see above) have been shown to be sensitive to and negatively affected by
saltwater intrusion over the decade-long course of the study. This study suggests a
biologically plausible link between sulfide increases caused by SOx deposition and tidal
marsh intrusion upon tidal swamp zones. However, the authors did not consider sulfate as
originating from any source other than seawater, so the effect of salinity could not be
separated from the effect of sulfate reduction and resultant sulfide phytotoxicity. There is
no evidence of SOx deposition contributing to sulfide phytotoxicity in wetland and
aquatic ecosystems with significant seawater contributions.
Sulfide toxicity has not been documented for freshwater animal species, since the
reducing conditions that produce sulfide also cause hypoxia. However, animal species do
vary in their tolerance of sulfate concentrations in water. Carlisle et al. (2007) and
Meador and Carlisle (2007) used data from the U.S. Geological Survey's 1993-2003
National Water-Quality Assessment (NAWQA) to generate indicator values for widely
distributed North American macroinvertebrates and fish. Each indicator value is the
average value for a water quality parameter at which a taxon is detected in NAWQA
samples. A follow-up analysis of macroinvertebrates and stream quality in the eastern
and midwestern U.S. designated streams with more than 20% loss of expected taxa for
the region as degraded habitat (Carlisle and Meador. 2007). This dataset has identified
animals sensitive and tolerant to sulfate and other water quality parameters, and allows
the evaluation of sulfate impacts on animal species at concentrations lower than those
that cause acidification.
Internal Eutrophication
Internal eutrophication occurs in saturated soils in freshwater wetlands and lakes when
sulfate addition results in the release of bioavailable phosphate from soil forms that are
not bioavailable. Additional sulfate in surface waters can increase P mobility and cause
12-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
internal eutrophication of the recipient ecosystem; farmers growing sugarcane in the
oxidized peatland soils of the Florida EAA use S as a soil amendment for this reason (Ye
et al.. 2010).
Wetland and aquatic systems often retain phosphate as FePC>4. When sulfate is reduced,
the resultant sulfide binds with iron to create FeS, which precipitates out of solution,
releasing P into surface waters (Figure 12-4). This internal eutrophication can alter plant
community composition (U.S. EPA. 2008a). or can speed the accretion of organic matter
and infilling of the wetland, reducing wetland habitat (Kleeberg et al.. 2016). In
mesocosms resembling shallow lakes planted with wild rice, experimentally elevated
surface water sulfate concentrations raised phosphate concentrations in the water column
(Pastor etal.. 2017). Water and sediment mesocosms of samples collected from Lake
Moshui in China showed that sodium sulfate addition to raise sulfate concentrations to
500 mg/L also raised peak [P] by 180% in the water column, and by 210% in the
sediment pore water, above unamended control mesocosms, and the peak [P] with sulfate
addition occurred a week later than in control mesocosms (Yu et al.. 2015). Limitation of
microbial sulfate reduction achieved through experimental NOs enrichment of the
hypolimnion in Onondaga Lake, NY, decreased soluble reactive phosphorus in the
hypolimnion by 95% (Matthews et al.. 2013). Sulfate addition can alter P dynamics and
cause eutrophication in freshwater systems.
12-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Fe(ll) = iron (II); Fe(lll) = iron (III); FeSx = iron-sulfide complexes; P043 = phosphate.
Open boxes indicate pore water components, and shaded boxes indicate solid-phase chemistry. Reduction reactions are indicated
by solid arrows, oxidation reactions are indicated by dotted lines, and phase changes (precipitation or dissolution) are indicated by
dashed lines.
Source: from Simkin et al. (2013).
Figure 12-4 Mechanisms of linked sulfur, iron, and phosphorus cycling in
wetland waters and soils.
Surface water sulfate may decrease heavy metal bioaccumulation in plants, but the
implications of this process for impacts of SOx deposition may be limited. Sulfate
addition can decrease the accumulation of toxic elements in aquatic plants. In soil, S
addition decreased dissolved arsenic (As) concentrations in soil solution, but increased
As(III):As(V), although the researchers suggested that the formation of As-Fe-sulfide
minerals prevented plant As uptake (Jia et al.. 2015). In microbially spiked soil
mesocosms, the addition of SRB had a stronger stimulatory effect on As(III) As(V) than
did S addition. In rice mesocosms, S addition decreased As concentrations in rice roots
and in the iron plaques on roots (Jia et al.. 2015). In laboratory mesocosms, accumulation
of selenium (Se) in aquatic plants was a function of both sulfate and Se concentrations in
water, but increasing sulfate concentrations decreased Se concentrations in both Lemnct
minor (aquatic macrophyte) and Pseiidokrichneriella subcapitata [unicellular green alga
(Lo et al.. 2015)1. Sulfate in surface waters may slow bioaccumulation of heavy metals in
biota by forming relatively insoluble S-metal complexes and preventing metal uptake by
aquatic plants. However, these findings may have limited applicability to SOx deposition.
If there is a common source of both SOx and heavy metal deposition to an aquatic
environment, it is unlikely that deposition will correlate with decreasing metal
bioavailability. Also, mercury is a focus of research into heavy metal environmental
transformations and bioavailability, and SOx deposition is linked instead to increased
July 2018
12-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
bioaccumulation of Hg (see Appendix 12.3 to Appendix 12.6). There is no research on
how SOx deposition affects Se or As accumulation in plants growing under field
conditions.
Sulfur Effects on Methane Emissions
The 2008 ISA documented the suppression of methane emissions from aquatic and
wetland ecosystems in response to increases in sulfate concentrations. When sulfate is
added to freshwater systems, SRPs outcompete methanogens, and methane emissions are
suppressed. The activity and environmental preferences of SRPs are key to understanding
the effects of SOx deposition upon methane and MeHg production. Under anaerobic
conditions, SRPs and methanogens compete for organic C as a metabolic substrate, and
the reactions of sulfate reduction coupled with organic C oxidation are
thermodynamically favored over the reactions of methanogenesis (Paulo et al.. 2015). In
many anaerobic environments, sulfate is in short supply, and methanogens are more
active than SRPs in mineralizing organic C. Sulfate addition may swing the competitive
balance back towards SRPs, decreasing methane production in saturated soils and surface
waters. However, studies summarized in other parts of this appendix show that SRPs and
methanogens can form syntrophies (see Appendix 12.3.2) or coexist and rapidly shift
dominance within a single wetland (see Appendix 12.3.2 and Appendix 12.3.3.1).
The relative primacy of methanogens versus SRB in oxidizing carbon is one of the
factors which distinguishes freshwater from coastal wetlands and water bodies. Older
literature suggests that when SOx deposition is not a factor, methanogens dominate
anaerobic decomposition in freshwater systems (including wetlands), with
methanogenesis responsible for 72-82% of organic C mineralization at the ELA, Ontario
(Kelly and Rudd. 1984). and methanogenesis responsible for 4 times the electron flow
and organic C mineralization of sulfate reduction in Lake Mendota, WI (Ingvorsen and
Brock. 1982). In estuarine and marine ecosystems, on the other hand, SRPs are the
dominant anaerobic decomposers. In the saltwater marshes at Kirkpatrick Marsh, MD,
methanogenesis is <10% of C mineralization, and sulfate reduction is responsible for
45-85% of total C mineralization in the summer, and 5-25% of C mineralization in the
fall (Mitchell and Gilmour. 2008). Older research suggests that in pristine ecosystems,
methanogens dominate anaerobic C mineralization in freshwater systems.
New research on microbial communities in saturated sediments contributes to the
biological plausibility of the reduction in methanogenesis caused by SOx deposition
established by the 2008 ISA . Adding sulfate to freshwater systems can depress
methanogenesis because anaerobic C mineralization by SRPs is energetically favored
12-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
when sulfate and acetate are available at the sediment surface (Urban etal.. 1994). Newer
work also supports this finding (see Table 12-3). Twitchell Island is a restored wetland in
the San Joaquin delta, CA, where researchers sampled sediments and water along a
gradient of decreasing riverine S load (14-8 mg sulfate/L in February) from the wetland
inlet to a transitional site, and to an interior marsh site. In summer, sulfate was 60% lower
and methane emissions were 110% higher at the interior site than at the inlet, and 16S
rRNA sequencing confirmed that relative SRP abundance was negatively correlated with
methanogen abundance (He et al.. 2015). In the Everglades WCA, FL, similar gradients
of S loading occur from agricultural runoff in the freshwater marsh. Sampling at sites F4
(7.1 mg/L or 74 (j,M sulfate in pore water), U3 (3.7 mg/L or 39 (j,M sulfate), and W3
(<0.4 mg/L or <4 (j,M sulfate) found effects of sulfate upon SRB (sulfur-reducing archaea
were not sampled) and methanogens (Bae et al.. 2015). SRB abundance increased with
increasing sulfate pore water concentrations across the three sites, and was 6.9 times
higher at U3 and 16.8 times higher at F4 than at W3. Sulfate also affected the competitive
interaction between SRB and methanogens, which had equal abundances at the most
pristine W3 site. At the sites with higher S loads, SRB were 60-80% more abundant than
methanogens. The relative abundance of methanogens was determined by copy number
of the mcrA gene in environmental samples, which correlated positively with methane
production rates and mRNA (Bae et al.. 2015). so sulfate loading alterations to the
microbial community in the Everglades WCA may also be lowering methane emissions.
Sampling from natural microbial communities confirms that sulfate favors SRB over
methanogens.
12-19
DRAFT: Do Not Cite or Quote

-------
Table 12-3 New studies on nonacidifying sulfur effects on methane emissions.
Type of
Ecosystem
Additions or
Load (kg
S/ha/yr)
Biological and Chemical
Effects
Study Site Study Species
Reference
Restored River water is
peatland the source of
island in the S load,
river	more interior
sites have lower
S load, as
reflected by
February
sediment
surface
measurements:
Inlet [SO42"]
14 mg/L;
Transitional
[S0421
11.5 mg/L;
Interior [SO42"]
8 mg/L
SRP abundance decreases Twitchell
with increasing
methanogen abundance
(r= - 0.67).
Denitrifier abundance
decreases with increasing
methanogen abundance
(r= -0.64).
In August, sulfate
decreases 60% from inlet
to interior, and methane
emissions increases 110%.
SRP relative abundance
increases with sulfate
concentration (r= 0.96).
Island, San
Joaquin
delta, CA
Island covered He et al. (2015)
with Typhus spp.
and
Schoenoplectus
acutus, sampled
at inlet,
transitional, and
interior marsh
sites.
Microbial
community
quantified by 16S
rRNA sequencing
of peat sample.
Freshwater
marsh
W3 pore water
[SO42"] <4 |jM
U3 pore water
[SO42"] is 39 |jM
F4 pore water
[SO42"] is 74 |jM
Deposition = not
measured
SRB abundance is 6.9x
higher in U3 than W3 and
16.8x higher in F4.
In W3, abundance of SRB
and methanogens are not
significantly different. In
U3, SRB abundance is
80% higher than
methanogen abundance. In
F4, SRB abundance is
60% higher than
methanogens.
mcrA copy number
correlates positively with
mRNA and methane
production rates.
Acetotrophic methanogens
are dominant at W3, while
at high S U3 and F4 sites,
hydrogenotrophic
methanogens are
dominant.
F4, U3, and
W3 sites in
Water
Conservation
Area,
Everglades,
FL
Methanogens as
quantified by mcrA
copies.
Sulfur-reducing
bacteria as
quantified by dsrB
copies.
Bae et al. (2015)
July 2018
12-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Table 12-3 (Continued): New studies on nonacidifying sulfur effects on methane
emissions.
Type of
Ecosystem
Additions or
Load (kg
S/ha/yr)
Biological and Chemical
Effects
Study Site Study Species
Reference
Coastal salt
marsh
Total load = not
measured
Deposition = not
measured
No significant correlation
between potential methane
emissions and pore water
S042", NO3-, and Fe3+.
Methanogen abundance is
84% lower in S. alterniflora
than in P. australis marsh.
SRP abundance is 73%
lower in S. alterniflora than
in P. australis marsh.
Methanogen abundance
(M, as 1,000 gene copies/g
soil) increases with pore
water NO3" concentrations
(Nbad, as |jM NO3"):
M = 0.3468 + 0.8149 x
Nload
Shanyutan,
Min River
Estuary,
China
Zones of Chinese
native Phragmites
australis, Chinese
invasive
(American native)
Spartina
alterniflora, and
Cyperus
malaccensis
(subtropical
species not
present in U.S.)
Methanogenic
archaea quantified
using 16S rRNA
SRPs are
quantified using
dsrA gene
Tona etal. (2015)
16S rRNA = 16S ribosomal ribonucleic acid; Fe3+ = iron; g = gram; ha = hectare; kg = kilogram; L = liter; |jM = micromolar;
mg = milligram; N03" = nitrate; S = sulfur; S042" = sulfate; SRPs = sulfur-reducing prokaryotes; yr = year.
A laboratory study with controlled redox conditions and microbial communities suggests
that even when SRB and methanogens are not in direct competition for anaerobic C,
methanogen activity will increase when sulfate concentrations decrease. In an incubation
of water and sediment collected from oligotrophic Lake Stechlin, Germany, researchers
were able to create fluctuating redox conditions while continuously sampling the water
column. When the system was reduced, expression of the dsrB gene (present in SRB)
increased, and sulfide concentrations increased, indicating that SRB abundance and
activity increased in response to anoxic conditions in the water column (Frindte et al..
2015). Methane also increased in the water column, indicating no competition between
SRB and methanogens in this system. Over time, dsrB diversity decreased, and
expression ofpmoA (present in methanogens) increased (Frindte et al.. 2015). indicating
that although there was no competition between SRB and methanogens, methanogens
were able to persist after SRB had exhausted the available sulfate and diminished in
abundance. Active methanogenesis occurs when SRB activity is constrained by low
sulfate availability. Reductions in SOx deposition could increase methanogen activity
typical of wetland and aquatic sediments.
As with sulfide phytotoxicity, impacts of SOx deposition upon methanogenesis may be
present in freshwater but not marine ecosystems, as the high endogenous concentration of
sulfate in seawater is likely to overwhelm the impact of anthropogenic S. In marine
July 2018
12-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
sediments, sulfate reduction accounts for 50-100% of anaerobic carbon mineralization
(Urban ct al.. 1994); methanogenesis often persists at depths in organic sediments below
the depth to which sulfate can diffuse. In estuarine wetlands in the Cape Fear River
system in North Carolina, Hackney and Avery (2015) stated that in vertical soil profiles,
a concentration of 28 mg/L (300 (j,M) sulfate was the threshold where S reduction
switched to methanogenesis as the dominant form of C mineralization. This study also
determined that in estuaries, the transition from methanogenic to S-reducing conditions
occurs when more than 25% of tidal flooding carries more than 1 ppt of seawater
(Hackney and Avery. 2015). In a study of the Min River Estuary in China, there was no
relationship between methane emissions and pore water sulfate, but dominant plant
species did affect microbial communities, with lower relative abundance of both
methanogens (84% decrease) and SRPs (73% decrease) in Spartina alterniflora
sediments than in Phragmites australis-associated sediments (Tong et al.. 2015). Both of
these species are dominant plants in North American eastern coastal marshes, and this
result suggests plant community may be more important than sulfate deposition in
determining methane emissions in coastal marshes. SOx deposition is unlikely to affect
methane emissions from coastal marshes and water bodies. Sulfate addition may also
result in lower methane emissions by stimulating microbial methane oxidation. In marine
sediments, this occurs in syntrophic associations (associations among heterotrophs
dependent on the metabolic byproducts of other microbes) between anaerobic
methanotrophic archaea and SRB, which pair anaerobic oxidation of methane with sulfate
reduction, or in a clade (monophyletic group) of anaerobic methanotrophs capable of
both sulfate reduction and methanotrophy (Jove. 2012). However, these processes have
been observed only in marine deep-sea sediments, and it is not yet clear how widely
distributed or important they are in other systems. In peat bog mesocosms, by contrast, S
addition of 96 mg/L decreased microbial biomass 23-57% and decreased methane
oxidation to almost zero (Lozanovska et al.. 2016).
Interactions between Sulfur Deposition and Mercury
The 2008 ISA found the evidence was sufficient to infer a causal relationship between S
deposition and increased methylation of Hg, in aquatic environments where the value of
other factors was within adequate range for methylation. Recent research, as well as key
early papers on the microbial methylation of mercury, is summarized in this section. New
evidence is consistent with the 2008 ISA in identifying sulfur-reducing bacteria as a
biotic link between increased SOx deposition and increased MeHg concentrations in the
environment and in biota. New developments since 2008 include identification of both
12-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
the microbial genes linked to mercury methylation and the methylation capability in
certain archaeal strains (Appendix 12.3.1 and Appendix 12.3.2V
Ecosystem areas with high MeHg fractions (as a fraction of total Hg) are considered to
indicate active microbial methylation, as well as places where primary producers and
animals are at an elevated risk of MeHg accumulation. An active area of research is in
identifying areas or zones of ecosystems where %MeHg is elevated . Background
information on deposition and biogeochemical cycling of Hg, with particular emphasis on
identifying hotspots of mercury methylation, is presented in Supplemental Material.
Mercury methylation is a microbial process carried out by a subset of prokaryotes that are
active under reducing environmental conditions (anoxic water or sediments) while
requiring oxidized sulfur and organic C as substrates (see Figure 12-5). As a result, Hg
methylation rates are heterogeneous across time and the landscape. Transformation by
bacteria and archaea of inorganic Hg to organic MeHg occurs in wetland soils, in
Sphagnum mats of wetlands, and at the oxic-anoxic boundary in lakes, estuaries, and the
Arctic water column. Hg methylation by SRB embedded in periphyton is a new topic that
was not covered in the 2008 ISA (Appendix 12.3.2V Evidence of MeHg production in
periphyton has important implications for aquatic MeHg concentrations, because
periphyton may extend throughout water columns and are more connected to aquatic and
terrestrial food chains than are Hg methylation hotspots in sediment or peat. Regardless
of their location within the ecosystem, microbial Hg methylators have specific
environmental requirements which affect their rates of Hg methylation
(Appendix 12.3.3). There is new evidence from sulfur addition studies (Appendix 12.3.4)
and from studies of ambient conditions in North American ecosystems (Appendix 12.3.5)
to show that S deposition and ambient sulfate concentrations increase mercury
methylation under certain conditions. New evidence is consistent and coherent with the
conclusions of the 2008 ISA and the body of evidence is sufficient to infer a causal
relationship between S deposition and the alteration of Hg methylation in surface
water, sediment, and soils in wetland and freshwater ecosystems.
12-23
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
Sulfate-reducing prokaryote
Sulfate | Ł
Dissolved
Organic
Carbon
i
increases
availability
~r
Inorganic
Mercury
dissimilatory
sulfur reduction
I
metabolism
of carbon
methylation
by hgcAB
proteins
Sulfide
Increased
SRP Activity
or Abundance
Methyl
Mercury
Figure 12-5.
Sulfate, dissolved organic carbon, and inorganic mercury are all
important determinants of the rate and amount of methyl mercury
produced by sulfate-reducing prokaryotes.
12.3.1. Mercury Cycle and the Importance of Methylation
Mercury is present in ecosystems as elemental Hg", as Hg2+ complexed to metals or
organic compounds, or as MeHg. Since the Industrial Revolution, Hg emissions have
increased Hg concentrations in water and soils above preindustrial levels. Although in
recent decades U.S. emissions have declined [79% decrease in Hg emission between
1990 and 2011 as reported by the NEI (U.S. EPA. 2016a)l. Hg concentrations remain
elevated in air, water, soil, and biota above inferred preindustrial values. Environmental
mercury sources and transformations are illustrated in Figure 12-6; important mercury
sources to ecosystems that this appendix will consider include geological formations, wet
and dry deposition of Hg, and legacy Hg stored in soils from historical deposition or
historical (i.e., no longer active) industrial use. Hg released from active industrial
sources, active mines, or historical mine sites to soil or water are not considered.
July 2018
12-24
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Source: from Biaham et al. (2017).
Figure 12-6 Cycling of mercury (Hg) and methylmercury (MeHg) in
ecosystems.
Hg deposition is generally more bioavailable and readily methylated than is Hg already
stored in sediments or organic matter within the ecosystem, and sulfate stimulation of
methylation favors Hg newly added to the system over Hg already in the ecosystem
(Bigham et al.. 2017).
hi terrestrial soils, Hg forms complexes with organic matter (Poulin et al.. 2016). New
research suggests that disturbances that accelerate carbon fluxes in terrestrial systems
also accelerate Hg release from soils and sometimes stimulate methylation. In a study of
Picea cibies (Norway spruce) stands in Sweden, clear-cutting raised the soil water table
and created conditions favorable for Hg methylation. Clear-cutting did not alter Hg
concentrations in the organic soil (O) layer, but it did significantly increase MeHg
concentrations and %MeHg in the O layer 5-7 times over control conditions (Kronberg
et al.. 2016). hi a similar study of clear-cutting plantations planted on drained peat soils in
Finland, harvesting increased DOC and Hg in drainage waters (Ukonmaanaho et al..
2016). Peat SRB abundance correlated with MeHg concentrations and varied by geology,
with higher MeHg concentrations in watersheds with underlying schist formations (high
Hg content). In Oak Ridge, TN, riparian forest soils along East Fork Poplar Creek are
contaminated by historical industrial mercury releases and contain 2-3 times the Hg in
uncontaminated soils. Soil microcosms showed that flooding of these soils moved
mercury bound to soil organic matter into the soil solution, as aqueous Hg" or as HgS
precipitates, the latter produced by microbial sulfate reduction within the soils (Poulin et
al.. 2016). Both Hg" and nanoparticulate HgS are bioavailable to microbial methylators,
My 2018
12-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
and flooding resulted in a pulse of both aromatic DOC and MeHg in the riparian soil pore
waters (Poulin et al.. 2016).
Recent studies using stable isotopes of mercury reviewed by Paraniape et al. (2017)
reported a range of 0.1-15% MeHg in North American and Swedish wetlands, and
0.4-4.6% MeHg in lake sediments. There is new evidence that MeHg concentration in
Adirondack surface waters is positively correlated with Hg° deposition rate (Gcrson and
Driscoll. 2016). Specifically, Hg in terrestrial leaf litter from Arbutus Lake watershed
was measured 6 times over 2004-2014, and declined 40% in concert with a 25% decline
in Hg° atmospheric concentrations. Over the same time period, MeHg concentrations in
Arbutus Lake declined (p < 0.03), although there was no quantification of the trend
reported. Gerson and Driscoll (2016) suggested that these results show that decreases in
Hg° concentrations will limit the supply of bioavailable Hg to mercury methylators and
decrease MeHg in aquatic ecosystems.
Wetlands tend to have higher MeHg fractions than other water bodies at similar locations
and order within watersheds. In the Adirondack Mountains, NY, sampling of soils and
substrates (upland, riparian, open water, peat bog) at different points in the Sunday Lake
watershed found that the highest MeHg fraction was in Sphagnum mats (Yu et al.. 2010).
In the Archer Creek watershed, also in the Adirondacks, MeHg concentrations in
wetland-draining streams were elevated compared to MeHg in upland streams
(Selvendiran et al.. 2008a). At a larger scale survey of 44 Adirondack lakes sampled
2003-2004, land cover did not correlate with MeHg in lake water or biota, possibly
because variation in wetland cover was low across watersheds (Yu et al.. 2011). In a
survey that included 21 freshwater, brackish, and marine wetlands and lakes in the
Mississippi River delta near Lake Pontchartrain, LA, MeHg levels in wetlands were
higher than in nearby rivers or lakes with similar salinity levels. In brackish and
freshwater wetlands, total MeHg concentrations (ng/L) and MeHg fraction (MeHg/total
Hg) were 2 or 3 times as much as in brackish or freshwater rivers (Hall et al.. 2008).
The relationship between Hg and MeHg concentrations may depend heavily on the type
and salinity of water body. A recent report on Hg concentrations in streams across the
U.S. found no relationship between Hg and MeHg across sites (Wentz et al.. 2014). This
is in contrast to an earlier review of Hg methylation and demethylation dynamics found
that across studies, there is a positive correlation between Hg and MeHg in estuaries
(R2 = 0.78), and weaker positive correlations in rivers (If = 0.68) and lakes (If = 0.64),
but no correlation in wetlands (Benoit et al.. 2003). In a recent study of eight highly
Hg-contaminated (up to 226,000 ng Hg/g sediment, reported as 226 |ig Hg/g sediment)
lakes and water bodies in Sweden, more than 55% of the variation in sediment MeHg was
explained by total sediment Hg in three estuary sites, but there was no correlation in the
12-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
remaining freshwater sites (Drott et al.. 2008). In the freshwater sites, MeHg was instead
strongly correlated with potential methylation rates [r1 > 0.85 (Drott et al.. 2008)1.
Biology of Sulfate-Reducing Prokaryotes
The 2008 ISA attributed Hg methylation to sulfur-reducing bacteria, and there is new
evidence of the link between microbial sulfate reduction and Hg methylation. New (and
key older) studies include experimental inhibition of different microbial groups in order
to demonstrate microbial mechanisms, molecular sequencing techniques to identify
potential microbial Hg methylating strains, and sequencing of environmental or
experimental samples to determine microbial dynamics of Hg methylation. These studies
were conducted with samples from rivers, lakes, wetlands, and laboratory mesocosms
(Table 124). Current knowledge suggests mercury is methylated in the environment by
certain strains of sulfur-reducing and iron-reducing bacteria in Deltaproteobacteria
(Desulfovibrio, Desulfotomaculum, Desulfobulbus, and Geobacter genera) as well as
methanogens in Archaea, while the genetic potential for methylation has been identified
in the bacterial Firmicutes (Paranjapc et al.. 2017).
Increasing SOx in the microbial environment increases the metabolic activity and growth
of sulfate reducers, and one of the metabolic activities that can be stimulated is the
methylation of Hg. Peat samples from the Bog Lake fen in the Macell Experimental
Forest showed that experimental increases of wet S deposition to historical levels (32 kg
S/ha/yr) changed entire bacterial community composition. Experimental S addition also
significantly changed the composition of deltaproteobacterial communities, which
contain the SRB and IRB, and this community change correlated with bog %MeHg
(Strickman et al.. 2016). In recovery sections of the same wetland (S treatment of 32 kg
S/ha/yr had ceased 3 years before sampling), entire bacterial community and
deltaproteobacterial community compositions were not significantly different from
control wetland communities (Strickman et al.. 2016). suggesting a rapid response time
of microbial communities to changes in SOx deposition.
12-27
DRAFT: Do Not Cite or Quote

-------
Table 12-4 New studies on the biology of sulfate-reducing prokaryotes.
Additions or
Type of	Load Biological and Chemical
Ecosystem (kg S/ha/yr)	Effects	Study Site Study Species	Reference
Peatland
Treatment:
Bacterial communities in
Marcell
Pore water. Strickman et al.

32 kg
areas with increased
Experimental
bacterial (2016)

SO42 /ha/yr
sulfur deposition had
Forest, MN
communities

addition,
significantly different



control/ambient:
compositions compared



4.6 SO42"
to control and recovery



kg/ha/yr
areas. Bacterial species




diversity was significantly




higher in the control than




sulfate treated or recovery




groups.




Deltaprotobacteria were




significantly related to




%MeHg in bogs and close




to significant in lagg areas




(p = 0.057).


River
Deposition not
Sulfate reduction rates
Mesocosms
Avramescu et al.

reported
were positively correlated
of sediments
(2011)


with potential Hg
from St.



methylation rates
Lawrence



(r2 = 0.98).
River



Methane production in




sediment slurries




correlated positively with




demethylation rates




(r2 = 0.61).




Direct inhibition of




methanogens decreased




demethylation by 83%.


Floating
2.0 mM
Addition of M0O42"
Sunday Lake,
SDhaanum sdd. Yu et al. (2010)
bog
(190 mg/L)
decreased MeHg
Adirondack


sulfate added to
production by 44%.
Mountains,


slurried
Potential Hg mthylation
NY


samples
rates (%MeHg/day) were




2.1x higher with SO4




amendment.


Lakes
Deposition not
Addition of M0O42"
Lake
Samples of Gascon Diez et al.

reported
decreased methylation by
Geneva,
Darticles from (2016)


60-90% in settling
Switzerland
water column and


particles, and by 80% in

from lake bottom


sediments.

sediments
July 2018
12-28
DRAFT: Do Not Cite or Quote

-------
Table 12-4 (Continued): New studies on the biology of sulfate-reducing
prokaryotes.
Type of
Ecosystem
Additions or
Load
(kg S/ha/yr)
Biological and Chemical
Effects
Study Site Study Species
Reference
Lakes
Deposition not
reported
Addition of M0O42"
decreased MeHg
production by 30% for E.
crassipes and by 60% for
P. glabra.
qPCR of bacterial genes
indicated that 3.34% of
the total bacterial
community in the
periphyton belonged to
the SRB families
desulfovibrionaceae and
desulfobacteraceae.
MeHg fraction was
positively correlated with
relative abundance of
desulfobacteraceae in P.
glabra periphyton.
Amazon
Oxbow
Lakes,
Bolivia
Eichhornia
crassipes and
polygonum
densiflorum
(currently named
persicaria glabra)
Acha etal. (2011'
Freshwater
marsh
W3 pore water
[S042"] <4 |jM
U3 pore water
[SO4"] is 39 |jM
F4 pore water
[SO42"] is 74 |jM
Deposition = not
measured
SRB abundance is 6.9x
higher in U3 than W3 and
16.8x higher in F4.
In W3, abundance of SRB
and methanogens are not
significantly different. In
U3, SRB abundance is
80% higher than
methanogen abundance.
In F4, SRB abundance is
60% higher than
methanogens.
mcrA copy number
correlates positively with
mRNA and methane
production rates.
Acetotrophic
methanogens are
dominant at W3, while at
high S U3 and F4 sites,
hydrogenotrophic
methanogens are
dominant.
F4, U3, and
W3 sites in
Water
Conservation
Area,
Everglades,
FL
Methanogens as
quantified by mcrA
copies
Sulfur-reducing
bacteria as
quantified by dsrB
copies
Bae etal. (2015)
July 2018
12-29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Table 12-4 (Continued): New studies on the biology of sulfate-reducing
prokaryotes.
Additions or
Type of	Load
Ecosystem (kg S/ha/yr)
Biological and Chemical
Effects
Study Site Study Species
Reference
Marine Laboratory
Methylation increases	Laboratory
with increasing bacterial	cultures
biomass (r = 0.83).	Sediments
During exponential growth	from Berre
phase, S addition	Lagoon,
increases net MeHg	France
production by 50%.
hgcAB expression levels
do not correlate with
methylation capacities.
Desulfovibrio	Goni-Urriza et al.
dechloracetivorans, (2015)
strain BerOCI from
sediments
cultures with
different C
sources
(pyruvate,
fu ma rate,
lactate, ethanol,
or malate),
control and
additional
sulfate (30 mM)
C = carbon; ha = hectare; kg = kilogram; MeHg = methylmercury; mM = millimolar; |jM = micromolar; S = sulfur; S042 = sulfate;
SRB = sulfur-reducing bacteria; yr = year.
Sulfate-reducing bacteria (SRB) and iron-reducing bacteria (IRB) are primarily
responsible for Hg methylation in natural ecosystems. On a broad scale, these bacteria
metabolize organic C by pairing its oxidation with the reduction of sulfate and/or Fe3+.
The sulfide produced by the reduction of SOx can act as a negative feedback on SRP
activity and Hg methylation. Methanogens, SRB, and IRB are facultative anaerobes,
(Paraniape et al.. 2017). and their presence and activity within wetland and freshwater
ecosystems is limited to zones with low oxygen availability. Experimental inoculations
suggest that microbial methylators are capable of forming syntrophies, in which microbes
link their metabolisms by the exchange of metabolic substrates and products. This
suggests that sulfate stimulation of methylation may involve more types of organisms
than just the sulfate-reducing bacteria and has led to the classification of the group of
interest as sulfate reducing-prokaryotes.
Specific inhibition of SRPs using molybdate (M0O42 ) is a common tool in studies that
determine the contribution of SRPs to Hg methylation. An older study of acidic Dickie
Lake in Ontario used different microbial inhibitors to determine whether SRPs or
methanogens within the sediments were responsible for Hg methylation. Molybdate
inhibited sulfate reduction and decreased %MeHg by 75% of production in controls,
while 2-bromo-ethanesulfonate (BESA), an inhibitor of methanogens, did not
significantly decrease MeHg production (Kerry et al.. 1991). Researchers inferred that
SRPs, not methanogens, were responsible for methylation in the lake, and sulfate
reduction rates increased 0.35 mg sulfate reduced/L/day with a 1 mg/L increase in sulfate
addition. However, additions of sulfate to concentrations of 5-25 mg sulfate/L in
sediment slurries did not affect %MeHg, which was 0.20-0.25% MeHg/g sediment
(Kerry et al.. 1991). A similar relationship between MeHg and sulfate reduction was
July 2018
12-30
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
demonstrated using molybdate inhibition in coastal Georgia marsh soils, where Mo
addition decreased sulfate reduction by 90% and decreased MeHg production by 85%
(King et al.. 1999). indicating that SRPs were involved in both processes.
An early laboratory study cultivated strains of a SRB, a methanogen, or an acetogen with
an Hg spike. The SRBs were the only microbes that methylated added Hg, and they did
so at a rate that resembled the rate in aquatic sediments, while molybdate inhibited MeHg
production and demethylation of added MeHg (Pak and Bartha. 1998). In a more recent
study, incubations of St. Lawrence River sediments with different inhibitors showed that
molybdate inhibited sulfate reduction completely, but did not affect methane production
(Avramcscu et al.. 2011). indicating that methanogens did not play an important role in
reducing S. Moreover, sulfate reduction rates were strongly positively correlated with
potential Hg methylation rates (r2 = 0.98). However, when methanogens were inhibited,
methylation of Hg increased 16% (Avramcscu et al.. 2011). indicating that methanogens
compete with SRPs and can depress Hg methylation. In Everglades, FL sediment
incubations, molybdate decreased MeHg production by 90% (Gilmour et al.. 1998). In
incubations of Sphagnum sampled from Sunday Lake, NY, molybdate decreased MeHg
production by 44% (Yu et al.. 2010). In core incubations from lakes in Wisconsin,
including both basins of Little Rock Lake, molybdate reduced MeHg production by 50%
(Gilmour et al.. 1998).
In wetlands within the Allequash Creek watershed, WI, sediment incubations with added
14C-labelled substrate and sodium molybdate suggested a diverse variable microbial
methylation community (Crcswell et al.. 2017). In some incubations, addition of
molybdate increased 14CC>2 produced from the substrate, suggesting that iron-reducing
methylators were dominant; in some incubations molybdate decreased 14CC>2, suggesting
that sulfate-reducing methylators were dominant. Incubations from the lower part of the
wetland produced 14CH4 as their dominant product, suggesting that methanogens
dominated the microbial methylation at this location (Crcswell et al.. 2017). Dominant
microbial methylators varied seasonally and spatially, including by sediment depth.
In Lake Geneva, Switzerland, mercury methylation rates were an order of magnitude
higher in settling particles than in sediments, and molybdate inhibition suggested that
sulfate-reducing methylators were responsible for 60-90% of methylation in settling
particles, and 80% of methylation in sediments (Gascon Diez et al.. 2016). Gascon Diez
et al. (2016) suggest that this study may be relevant to the Great Lakes, which like Lake
Geneva, are large, deep, and experience algal blooms due to nutrient inputs.
New research suggests that SRPs are important methylators of mercury in disturbed
terrestrial soils as well. In a Swedish study of the effects of forest harvesting on mercury
dynamics, clear-cutting induced changes in mercury methylators as inferred from soil
12-31
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
incubations (Kronbcrg et al.. 2016). Additions of sulfate and iron enhanced potential
methylation rates in clear-cut but not reference forest soils, indicating that SRPs and
iron-reducing bacteria were both active methylators in disturbed soils with high water
tables (see Appendix 12.3.2V The role of SRPs was confirmed by inhibition of
methylation by molybdate in soils from clear-cut stands, and a role of methanogens in
methylating mercury was shown by BES inhibition of methylation (Kronbcrg et al..
2016).
The 2008 ISA did not address the role that periphyton play in hosting SRPs and boosting
Hg methylation rates within the oxic water column of aquatic and wetland ecosystems.
This section summarizes new (and key older) papers that demonstrate that SRB in
periphyton methylate Hg in South American lakes as well as North American lakes and
wetlands. Periphyton are ecologically important as they form the base of the aquatic food
chain for many invertebrates and vertebrates. Periphyton mats are biofilms of algae,
bacteria, fungi, microinvertebrates, organic detritus, and inorganic particles, embedded in
a polysaccharide matrix, typically attached to a substrate of sediments or submerged
macrophytes. The microbial communities embedded within the periphyton can be quite
complex and diverse, including autotrophs and heterotrophs, SRB and S oxidizing
bacteria, and methanogens; periphyton microbial composition and microbial activity
depend on site conditions (Correia et al.. 2012). The polysaccharide matrix slows
exchange with the water column while intense aerobic metabolism at the matrix surface
quickly consumes oxygen diffusing into the periphyton, creating anoxic, reducing
microenvironments in the interior of periphyton where anaerobes including SRPs thrive.
In the case of periphyton attached to macrophytes, plant exudates are an important source
of carbon for periphyton microbes, boosting activity above rates of activity of similar
microbes residing in sediments.
SRPs embedded in periphyton are more efficient in methylating Hg than are SRPs in
sediments. Much of the research demonstrating Hg methylation by periphyton has
occurred in South America, in tropical lakes where the genus Eichhornia (water
hyacinth) is native, and where ambient temperatures can support high levels of bacterially
mediated Hg methylation (Guimaraes et al.. 2006). Eichhornia crassipes is a wetland
plant present in 24 states and listed as a noxious or invasive weed by 7 states (USDA.
2015b). Closely related Eichhornia azurea is currently present only in Florida; it is listed
by the Animal and Plant Health Inspection Service of the U.S. Department of Agriculture
(USDA-APHIS) as a noxious weed (USDA. 2015b). In incubations of chopped
Eichhornia crassipes roots spiked with 203HgCl2, net MeHg production was high,
1.6-30.2% of total Hg (Guimaraes et al.. 2006). Other studies have shown a similarly
wide range of methylation by periphyton, with high rates exceeding sediment methylation
rates. Intact E. crassipes periphyton incubations had potential methylation of 12.1-25.2%
12-32
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
of added Hg, and periphyton isolated from Polygonum densiflorum [hereafter called by
its current accepted scientific name Persicaria glabra; a wetland obligate plant also
native to the U.S., classified as endangered in Maryland and New Jersey, (USDA.
2015b)! had potential methylation rates of 0.2-36.1%; by comparison, sediment from the
same tropical lakes produced 6.4-12.5% MeHg in incubations (Correia et al.. 2012).
Mo inhibition of sulfate reduction also inhibits MeHg production in periphyton,
confirming that SRPs are responsible for a significant portion of Hg methylation within
these complex microbial communities. Recently, Achaetal. (2011) and Correia etal.
(2012) collected macrophyte samples from oxbow lakes in the Bolivian Amazon for
inhibition experiments. Incubations with an array of guild-specific inhibitors show the
relative contribution of different microbes to Hg methylation. One set of incubations of
periphyton associated with the roots of E. crassipes or P. glabra suggested that SRB were
the most important component of the periphyton in producing MeHg. Molybdate, an
inhibitor of sulfate reduction, decreased MeHg production by 30% fori?, crassipes and
by 60% for P. glabra (Acha etal.. 2011). indicating that in intact periphyton, SRB alone
are responsible for 30-60% of Hg methylation. Quantitative PCR of bacterial genes
indicated that 3.34% of the total bacterial community in the periphyton belonged to the
SRB families Desulfovibrionaceae and Desulfobacteraceae, and MeHg fraction was
positively correlated with relative abundance of Desulfobacteraceae in P. glabra
periphyton (Achaetal.. 2011). A second set of incubations of six tropical macrophyte
species-associated periphyton confirmed that inhibition of sulfate reduction and SRB
using molybdate decreased MeHg production, but showed that inhibiting both SRB and
methanogens resulted in a 90% reduction of MeHg production across sediments and
periphyton (Correia etal.. 2012). Algaecide or fungicide additions significantly inhibited
MeHg formation in some but not all incubations, indicating that the microbial
composition or activity varied across periphyton samples, even when they were collected
from the same site or from the same host plant species grown at different sites (Correia et
al.. 2012). Together, these studies show that SRB are important in producing MeHg
within periphyton, but that other microbes residing in the periphyton contribute through
poorly understood mechanisms to total MeHg production.
Mercury methylation by periphyton has also been documented within North American
ecosystems. In an older study in Dickie Lake in Ontario, measured Hg concentrations in
periphyton were 56.5 ng/g, 22 times higher than Hg concentrations in sediments (Kerry et
al.. 1991). A more recent study of periphyton growing in the epilimnion of boreal Lake
Croche in Quebec demonstrated that Hg methylation rates are faster in periphyton,
reaching steady-state in 12 hours, than the 4-8 days to steady state of MeHg production
in sediments (Desrosiers etal.. 2006). MeHg production by periphyton in the boreal lake
was dependent on SRB as well as on photosynthesizers, as demonstrated by inhibition
12-33
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
experiments. Despite the lower temperatures of incubations to reflect the lower
temperatures of the boreal lake, potential MeHg production was 16.6-18.5% (Dcsrosicrs
ct al.. 2006). similar to MeHg production in tropical periphyton.
In the Florida Everglades, periphyton are the base of the aquatic food chain. In an older
study, incubation of periphyton collected from different sites along the eutrophic gradient
between the EAA to the north and Everglades National Park to the south found MeHg
production in periphyton from all sites, although rates varied widely (C'lcckncr ct al..
1999). MeHg potential production rates were highest in periphyton collected at the
northern, eutrophic site, and MeHg production correlated positively with both S reduction
and S oxidation by photosynthesizing purple S bacteria (Cleckner et al.. 1999).
Periphyton may be an important direct source of MeHg to the food chain in the Florida
Everglades.
The gene pair conferring the ability to methylate Hg has been identified only recently,
after the 2008 ISA. No ecological or evolutionary advantage of the ability to methylate
Hg has been established (Kcrin et al.. 2006; Benoit et al.. 2003). and it appears to be an
inadvertent transformation by a corrinoid-dependent protein produced by prokaryotes for
a different metabolic pathway (Gilmour etal.. 2013). Recent phylogenetics work
indicates that the gene pair hgcAB confers the ability to methylate Hg, and that the gene
pair and Hg methylation are present in some but not all species within the SRB,
iron-reducing bacteria, syntrophic sulfur reducers in the Syntrophobacterales,
methanogens in the Archaea, and bacteria in the Clostridia (Gilmour et al.. 2013). The
confirmed and predicted Hg methylators from this study occur across alkaline, neutral, or
acidic environments, and are mostly heterotrophs that use sulfate, iron, and CO2 as
terminal electron acceptors.
MeHg production rates vary among species, and even among strains within species (Shao
et al.. 2012). A recent study of hgcAB in the freshwater marshes of the Florida Everglades
showed that syntrophs dominated the microbial methylators (49-65% of sequences) and
also were the dominant sulfate reducers, as shown by dsrB mRNA quantification
(75-89%) of sediments sampled within the water conservation areas (Bae et al.. 2014). In
a recent study of methylation using a strain of Desulfovibrio dechloracetivorans,
expression of the hgcAB gene did not correlate with methylation capacities of pure
cultures, which supports the supposition that Hg methylation is not the main purpose of
the proteins encoded by hgcAB. However, during the exponential growth phase of the
cultures, addition of 2,900 mg/L (30 mM) sulfate increased net MeHg production by 50%
(Goni-Urriza et al.. 2015).
Demethylation is also an important microbial Hg transformation to consider in natural
ecosystems. The capacity to transform MeHg to inorganic Hg (elemental Hg or Hg2+) is
12-34
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
widely distributed and occurs by different mechanisms in different microbial groups.
Microbial demethylation by aerobic and facultative Hg-resistant microbes possessing the
mer operon break MeHg into methane and Hg° via reductive degradation, whereas
methanogens, SRPs, and other prokaryotes break MeHg into carbon dioxide, methane,
and Hg2+ via oxidative demethylation. A recent study of St. Lawrence River, Ontario,
Canada sediments showed that methane production in sediment slurries correlated
positively with demethylation rates (r2 = 0.61), and direct inhibition of methanogens
decreased demethylation by 83% (Avramcscu ct al.. 2011). An earlier study tested a pure
culture of the archaeon Methanococcus maripaludis and found that it demethylated added
MeHg despite no observed methylation potential (Pak and Bartha. 1998).
Microbial demethylation also occurs in the same prokaryotes that methylate Hg; a
laboratory incubation of pure cultures demonstrated that Desulfovibrio desulfuricans (an
SRB that methylates Hg) demethylated added MeHg at rates similar to those of incubated
aquatic sediments (Pak and Bartha. 1998). In the same study of sediments of the St.
Lawrence River, Ontario, inhibition of both SRPs and methanogens decreased
demethylation of Hg by 96% (Avramcscu et al.. 2011). indicating that while
methanogens are responsible for the bulk of demethylation in these sediments, SRPs are
involved in demethylation.
New evidence of the genetic basis of Hg methylation and the activity of microbial
methylators in the environment and in the laboratory is consistent with the 2008 ISA's
link between sulfate reduction and Hg methylation. Mercury methylation by SRP is
determined by the same suite of environmental factors that enhance SRP activity or
abundance: sulfate, organic matter, and anoxic-oxic conditions (Figure 12-7). Mercury
methylation occurs in wetlands, lakes, reservoirs, streams, and rivers. Algal blooms,
periphyton, stratification of the water column, and fluctuating water levels all create
favorable conditions for sulfate stimulation of mercury methylation.
12-35
DRAFT: Do Not Cite or Quote

-------
log S04, uM
Source: from Glimour (2011).
Figure 12-7 The effect of different environmental factors on the relationship
between sulfate and mercury methylation. Methylmercury (MeHg)
accumulation is minimal at low and high sulfate concentrations,
with an optimum near 100 |jM sulfate (blue line). High dissolved
organic matter (DOM) will increase the magnitude of MeHg
production across the range of sulfate concentrations (red line).
Sulfide produced by methylators will inhibit further methylation if
it accumulates in the aqueous environment where methylation
occurs, shifting the MeHg optimum left (purple line). However if
ecosystem chemistry (iron [Fe], reoxidation of sulfide, organic
matter [OM]) allows for rapid sequestration of sulfide to large
particles and sediments, the relaxation of the negative feedback
of sulfide to sulfur-reducing prokaryotes (SRPs) will shift the
MeHg optimum right (green line).
July 2018
12-36
DRAFT: Do Not Cite or Quote

-------
1	12.3.3. Environmental Drivers of Mercury Methylation Potential
2	The 2008 causality statement linked S to increased Hg methylation when other conditions
3	are favorable, with a particular emphasis on dissolved organic carbon (DOC) as a control
4	on Hg methylation (U.S. EPA. 2008a'). This section is an overview of factors that control
5	Hg methylation. In a review, Bigham et al. (2017) summarized drivers of mercury
6	methylation in terms of their effect on mercury availability to methylators and on
7	microbial activity associated with Hg methylation (Table 12-5).
Table 12-5 Environmental factors that affect mercury (Hg) methylation.
Geochemical or Bioavailability Microbial
Physical Parameter of Hg to	Methylator
Affects	Methylators	Activity	Notes
Sulfur	Increase or Increase
decrease
Bioavailability of inorganic Hg will increase or decrease
depending on Hg-S bonds and chemical species
formed.
Increased sulfate can stimulate activity of
sulfate-reducing prokaryotes, some which may be
capable of Hg methylation.
Organic carbon	Increase or Increase
decrease
Availability of inorganic Hg can decrease if Hg binds to
organic carbon molecules of a size or structure that
makes the molecules resistant to microbial degredation.
Availability of inorganic Hg can increase if Hg binds to
organic molecules that are mobile in water and
accessible to microbial degredation.
Increased organic carbon can stimulate microbial
activity by increasing electron donors to metabolism.
Iron
Decrease	Increase or Bioavailability of inorganic Hg can decrease as Fe
decrease	reduces and binds to Hg.
Fe can increase microbial methylation by binding to
sulfide and preventing the negative feedback of this
product of sulfate reduction upon its production, which
coincides with Hg methylation in some SRP. Also, Fe
can stimulate the action of iron-reducing microbial
methylators.
Fe can decrease microbial methlyation by shifting
microbial communities towards nonmethlyating
organisms.
PH
Increase or
decrease
Increase or
decrease
Bioavailability of inorganic Hg is reduced at high pH.
pH between 4.5 and 9 is optimal for microbial activity.
July 2018
12-37
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Table 12-5 (Continued): Environmental factors that affect mercury (Hg)
methylation.
Geochemical or
Physical Parameter
Affects
Bioavailability
of Hg to
Methylators
Microbial
Methylator
Activity
Notes
Temperature
Increase or
decrease
Increase or
decrease
Bioavailability of inorganic Hg can increase if
weathering of Hg-containing rocks or sediments, or
decomposition of older organic matter, increases with
higher temperatures.
Bioavailability of Hg can decrease if Hg volatilization to
the atmosphere increases.
Microbial methylators have optimal temperatures for
growth.
Oxygen availability
Increase
Decrease
Bioavailability of Hg in reduced chemical species is low.
Microbial methylators are anaerobes, methylating Hg
under reducing conditions.
Drying and wetting
cycles (drought,
seasonal variation in
water levels, etc.)
Increase
Increase
Fluctuating water levels can increase the release of Hg
from particles in sediment.
Fluctuating water levels can increase the reoxidation
and release of S from sediments as sulfate.
Adapted from Biaham et al. (2017)
Seasonal patterns of Hg methylation and transport of Hg in many systems can affect
MeHg accumulation. Also, a number of chemical constituents can stimulate or inhibit Hg
methylation. Given our understanding of the microbial mechanisms of Hg transport and
methylation, and the complexity and diversity of aquatic environments, chemical controls
on MeHg processes are not fully established. This section is meant to lend context to the
description of how S deposition affects methylation, but it is not a complete review of the
chemical controls on biological Hg methylation in natural environments.
12.3.3.1. Sulfur and Methylation Potential
Sulfate in soil pore water, surface water, or in sediment pore waters can stimulate
mercury methylation. Sulfate is utilized as a terminal electron acceptor in heterotrophic
metabolism by some classes of bacteria, and in the course of this dissimilatory sulfur
reduction, these bacteria transform sulfate to sulfide and expel the reduced sulfur into the
environment as a waste product. Sulfate-reducing bacteria can also form syntrophies,
symbiotic exchanges of metabolic substrates and products, with methanogens and other
archaea, so that sulfate stimulates metabolism even in organisms that do not utilize it
directly as an electron acceptor. The current scientific understanding is that MeHg is
produced as an accidental byproduct in both sulfate-reducing bacteria and syntrophic
archaea. Conditions optimal for these organisms and which favor their activity over
July 2018
12-38
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
competing anaerobic heterotrophs probably occur in short time frames and in shifting
locations in many ecosystems. Sulfate stimulation of mercury methylation will probably
occur at an annual time scale at very low or undetectable background levels, with
occasional peaks of microbial methylation (hotspots) when other controls (temperature,
organic matter, anoxic microsites, or hypoxic zones) on microbial metabolism reach their
optima. At temporal and spatial hotspots of mercury methylation, sulfate concentrations
may drop as sulfate is converted to sulfide, particularly in shallow water bodies with
longer water residence times. As a result, concentrations of sulfate in surface water may
not correlate with sulfate inputs to the ecosystem or with MeHg produced by sulfate
reduction, particularly in aquatic or wetland ecosystems where there are active SRPs.
The byproduct of microbial sulfate reduction, sulfide, is also an important regulator of
mercury methylation. Sulfide dissolved in surface water or pore water inhibits microbial
methylation, and interferes with plant nutrient uptake (see Appendix 12.2.3). Sulfide can
also produce a negative feedback on mercury methylation by binding with inorganic
mercury in the water column and precipitating into cinnabar in the sediment, lowering the
bioavailability of Hg to microbial methylators. However, in water bodies with a high
fraction of mineral sediments or dissolved minerals, particularly iron, iron will bind to
sulfide and precipitate out of solution, preventing the formation of a negative feedback to
sulfur reduction and mercury methylation. In water bodies with longer residence times,
sulfide concentrations in sediments may correlate with microbial sulfate reduction, and if
SRPs are the primary methylators in the system, with MeHg concentrations.
Laboratory studies support field research in demonstrating that under controlled
conditions, sulfate additions increased Hg methylation in environmental samples. The
2008 ISA reported evidence of Hg methylation in lake water and sediment samples in
response to experimental sulfate addition. In slurries of sediment collected from Quabbin
Reservoir, MA, adding 4.8 mg/L (reported as 50 (j,M) sulfate to the slurry (ambient lake
water sulfate = 5.8-7.7 mg/L or 60-80 (iM) increased potential MeHg production 150%,
adding 9.6 mg/L sulfate (100 (j,M) increased MeHg production 160%, and adding
19.2 mg/L sulfate (200 (j,M) sulfate increased MeHg 410% above ambient lake water
MeHg (Gilmour et al.. 1992). Slurries can overestimate the potential rates of MeHg
production, so this study also incubated intact sediment cores with sulfate concentrations
from 0.3-110 mg/L (reported as 3-1,140 (j,M). Peak potential MeHg production was at
11 mg/L sulfate (110 (j,M), where MeHg was 8.6 ng MeHg/g, a 200% increase from
MeHg production at 0.3 mg/L sulfate [3 (j,M (Gilmour et al.. 1992)1. In longer
incubations of sediment cores from two lakes in Wisconsin, including Little Rock Lake,
adding 5.8-23 mg/L sulfate (60-240 (iM) changed sediments from net demethylating
(-100 ng MeHg/m2/day or -0.1 |ig MeHg/m2/day) to net methylating
12-39
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
(5,500 ng MeHg/m2/day or 5.5 |ig MeHg/m2/day) environments (Gilmour and Ricdcl.
1995V
Since the 2008 ISA, several new studies have provided experimental evidence of sulfate
stimulation of Hg methylation in a broader range of aquatic freshwater environments
(Table 12-6). Laboratory experiments using samples from wetlands at Sunday Lake in the
Adirondack Mountains of New York were conducted to determine potential Hg
methylation rates. When slurried peat-and-water samples from a floating fen (bog mat
composed of Sphagnum spp. and ericaceous shrubs) received an addition of sulfate to
raise the concentration of sulfate by 190 mg/L (initial sulfate concentration not given),
Hg methylation rates (%MeHg/day) were 2.1 times higher than unamended samples (Yu
ct al.. 2010). This result is consistent with evidence from observational (Appendix 12.3.5)
and S addition studies (Appendix 12.3.4) in peat bogs that sulfate stimulates Hg
methylation in these wetland systems.
There are three new studies that demonstrate sulfate stimulation of water and sediment
Hg methylation in samples taken from rivers. Slurried sediments collected in 2010 from
the South River, VA, were spiked with sulfate to raise sulfate levels from average
ambient levels of 19.2 mg/L to 38.2 or 96.1 mg sulfate/L. Sulfate addition significantly
increased the potential methylation rate of Hg (%/day), with methylation rates
1.6-2.6 times higher in sulfate-spiked sediments than in sediments with ambient sulfate
concentrations (Yu et al.. 2012). There was a significant difference in methylation rates
between the low and high S addition treatments in only one of three sampled river
sediments; this sample showed methylation was higher in the sediment when sulfate
levels were 96.1 mg/L (Yu et al.. 2012). In microcosm slurries composed of water and
sediment samples from the Wupper River in Germany, higher MeHg in water correlated
weakly (If = 0.28, both variables natural log-transformed) with higher sulfate
concentrations, with mean pore water sulfate of 32.2 mg/L and range of 2-223 mg/L
(Frohne et al.. 2012). Higher total Hg concentrations in water correlated weakly
(If = 0.20) with higher sulfate, although DOC was a stronger predictor of total Hg
[i?2 = 0.53 (Frohne et al.. 2012)1. Tsui et al. (2008) constructed mesocosms to mimic leaf
decomposition in the hyporheic zone of Minnesota streams and rivers, where conditions
are hypothesized to be favorable for the SRB that methylate inorganic Hg. Water samples
were collected from seven streams or rivers in Minnesota, with sulfate levels at collection
of 4.0-26.2 mg sulfate-S/L (with an ultrapure water sample as a control, 0.0 mg S/L), and
incubated w ith Acer sp. (maple) leaves. After 66 days of decomposition, the total Hg
released from the decomposing litter into solution as dissolved Hg increased with initial
sulfate concentration (R2 = 0.541,/? = 0.038), and the fraction of total Hg that had been
transformed to MeHg increased in a power law dependence upon initial sulfate
concentration [i?2 = 0.572,p = 0.030, see Figure 12-8. (Tsui et al.. 2008)1.
12-40
DRAFT: Do Not Cite or Quote

-------
o>
O)
X
35
30
25 ¦
20 ¦
15 ¦
10 ¦
5
0
100
80
60
40
20
0
10	20
DSO, (mg S L"1)
30
cn
X
CD
%MeHg = methylmercury fraction; DHg = mercury dissolved in mesocosm water; DS042 = sulfate dissolved in mesocosm water;
L = liter; mg = milligram; ng = nanogram; S = sulfur; THg = total mercury.
Source: Figure 2A in Tsui et al. (2008).
Figure 12-8 The relationship between surface water sulfate and total mercury
or methylmercury fraction in river/leaf litter mesocosms.
This evidence of sulfate stimulation of Hg methylation in river ecosystems is consistent
with observational studies of positive correlations between sulfate and Hg methylation in
rivers and streams.
July 2018
12-41
DRAFT: Do Not Cite or Quote

-------
Table 12-6 New mesocosm or incubation studies on sulfur addition effects on
methylmercury.
Type of Additions or Load Biological and Chemical	Study
Ecosystem (kg S/ha/yr)	Effects	Study Site Species
Reference
River	S deposition not
reported
Soil:
2.0-2.6 mg S/g soil
[SO42"] pore water
in microcosm:
2.2-223 mg/L,
mean 32.0 mg/L
Total Hg and MeHg were
positively correlated with
sulfate concentrations:
ln(THg) = 2.956 + 0.697 x In
(S042"),
ln(MeHg) = 4.962 + 0.473 x In
(S042").
Wupper River,
Germany
Sediment
pore water
Frohne et al.
(2012)
Floating
bog
2.0 mM (190 mg/L)
sulfate added to
slurried samples
Potential Hg methylation rates
(%MeHg/day) were 2.1 *
higher with SO42"
amendment.
Sunday Lake,
Adirondack
Mountains, NY
Sphagnum
spp.
Yu et al. (2010)
River
Ambient
Sediment MeHg (ng/g)
South River,
In-channel
Yu et al. (2012)

concentration in
increased linearly with pore
VA
surface


water is 200 |jM
water sulfate (|jM)

sediments


(19.2 mg/L) sulfate
concentrations in the river,





MeHg = 0.61 x [SO42"] - 0.08.



Addition of 400 |jM
(38.4 mg/L) or
1,000 |jM
(96.1 mg/L) sulfate
in anoxic sediment
slurries
SO42" addition in slurries
increased potential
methylation rates by 1.6-2.6x.
Fluvial
zones
(streams
and rivers)
S deposition not
reported
Initial water SO42"
concentrations
(mg/L): range
4.0-26.2, mean
9.2, median 6.1
Following aquatic anoxic
decomposition of leaf litter,
total dissolved Hg increases
in decreasing proportion to
initial SO42" concentration.
Dissolved %MeHg increases
in decreasing proportion to
initial SO42" concentration.
Mesocosms
constructed
with samples
from Cedar
Creek LTER
(leaf litter) and
seven streams
or rivers
(water), MN
/Acersp. and
stream
microbial
communities
Tsui et al. (2008)
g = gram; ha = hectare; Hg = mercury; kg = kilogram; L = liter; MeHg = methylmercury; mg = milligram; mM = millimolar;
|jM = micromolar; ng = nanogram; S = sulfur; S042" = sulfate; THg = total mercury; yr = year.
July 2018
12-42
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
12.3.3.2. Total Mercury Concentration and Methylation Potential
The 2008 ISA did not present information on how total Hg concentration affects rates of
MeHg production, in part because there is no consensus across studies as to whether there
is a linear relationship between total Hg and MeHg in water or soil. In part this is because
most reported MeHg rates are potential rates determined by experimental incubations
with added Hg, and methylation rates have been shown to be dependent on initial
concentrations of Hg (King et al.. 1999; Gilmour et al.. 1992). Additionally, different
chemical forms of inorganic Hg vary in bioavailability and affect methylation rates
(Kucharzvk et al.. 2015; Graham et al.. 2012).
Two studies that sampled MeHg fractions at different locations within the U.S. provide
information at the landscape level about factors that increase probability of detrimental
effects upon wildlife due to Hg availability. There is a large variability across systems in
MeHg fraction. In a study of five forested watersheds distributed across the U.S. (in
Vermont, Wisconsin, Colorado, Georgia, and Puerto Rico), total Hg flux out of the
watershed in the streams was highest at Rio Icacos, Puerto Rico (54,400 ng Hg/m2/yr,
reported as 54.4 |ig Hg/m2/yr) and lowest at Allequash Creek, WI [250 ng Hg/m2/yr or
0.25 |ig Hg/m2/yr (Shanlev et al.. 2008)1. MeHg fraction was highest at Allequash Creek
(14.8 % of total Hg) and lowest at Rio Icacos (0.7%), indicating that watershed flux of
total Hg does not directly indicate risk to organisms (Shanlev et al.. 2008). In a study of
eight streams in Oregon, Wisconsin, and Florida, MeHg fractions and concentrations
(range: <0.01-17.8 ng/g sediment) in streambed sediments were higher in nonurban than
in urban streams (Marvin-Dipasquale et al.. 2009).
12.3.3.3. Temperature and Methylation Potential
The 2008 ISA did not describe temperature or seasonal effects on Hg methylation. This
section presents information from key older papers and recent research on temperature
effects in lakes, wetlands, and watersheds. Seasonal patterns of microbial methylation are
directly controlled by temperature effects on microbial metabolism, as well as by indirect
temperature effects, such as oxygen depletion and stratification and mixing of water
bodies.
As a microbial process, Hg methylation is temperature dependent. In Lake Clara, WI,
incubation at in situ temperatures of seasonally sampled sediment cores from a lake depth
of 10.5 m showed that Hg methylation increased slowly from near 0% in April to peaks
of 1% methylation in August and September, with temperature accounting for 30% of
methylation rate variation (Korthals and Winfrey. 1987). Demethylation was also
July 2018
12-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
measured in incubations and displayed slightly different seasonal variation; it increased
sharply from 1% in May to 4% in early July, then fell sharply to 1% in late July, staying
low for the rest of the sampling year. As a result, the methylation:demethylation ratio (a
predictor of MeHg concentrations in water) rose above 1 in late July through August, and
then again in late September (Korthals and Winfrey. 1987). Timing of methylation and
demethylation peaks were similar in sediments sampled from a lake depth of 1.0 m,
although gross methylation and demethylation rates were slightly higher (Korthals and
Winfrey. 1987). Sampling across the ELA in Ontario suggests that seasonal patterns of
methylation in lakes are due to seasonal shifts in oxygen availability, as locations and
depths within the lake where methylation was documented in the summer had no
methylation potential when resampled in the fall following turnover (Ecklcv and
Hint.e1ma.nn. 2006).
Fall mixing of stratified lakes can have significant effects on total MeHg in lake food
chains. In all years of sampling at Onondaga Lake, NY, there was a significant peak in
MeHg concentrations and in MeHg:total Hg ratios in the epilimnion in the fall following
turnover (Todorova et al.. 2014). In Little Rock Lake, WI, researchers inferred that
considerable demethylation occurred in the fall following mixing because high MeHg in
the summer water column did not accumulate in sediments and was not reflected in low
spring MeHg water column concentrations (Watras et al.. 2006). Short weather events
can also affect Hg exchange. During the fall of 2006, regular weekly sampling in
Onondaga Lake, NY, captured the effects of strong winds and heavy rains mixing the
stratified layers of the lake upon MeHg dynamics. Dissolved oxygen values dropped in
the sampled surface waters, and MeHg concentrations rose 23% to 0.236 ng/L in surface
waters, even as MeHg concentrations dropped in the hypolimnion (Todorova et al..
2014).
There are seasonal patterns of Hg methylation in wetlands as well. In peatlands of
Allequash Creek Watershed, WI, MeHg concentrations and MeHg fraction (percentage of
total Hg) peak in early fall (Creswell et al.. 2008). In the Adirondack Mountains, NY,
sampling at the inlet stream of Arbutus Lake showed a pulse of MeHg entering the lake
in summer months from the upstream wetlands (Gerson and Driscoll. 2016). Sampling of
the streams in the lake watershed showed no seasonal patterns of MeHg in streams
draining forest; but in streams draining wetlands, MeHg increased 300% during the
growing season compared to winter MeHg concentrations (Selvendiran et al.. 2008a'). In
freshwater marshes in the Yolo Bypass, CA, seasonal wetland export of MeHg occurs
during the winter as well as the summer. In the summer, newly methylated MeHg is
drawn into the root zone of the sediment by plant transpiration, and in winter, diffusion
releases MeHg from the sediment into the water column (Bachand et al.. 2014). Seasonal
changes have also been documented in the salt and freshwater marshes at Kirkpatrick
12-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Marsh on the Chesapeake Bay, MD, with higher MeHg concentrations and higher
methylation rates in summer than in fall (Mitchell and Gilmour. 2008). and temperature
effects have been experimentally demonstrated in Georgia coastal marshes, where the
methylation rate was 2 times higher at 25°C than at 4°C, and was 30 times higher at 35°C
than at 4°C (King ct al.. 1999). Seasonal patterns are even observed in the Florida
Everglades, where December potential methylation rates were lower than March or July
rates, although water temperatures at midday in December were 18°C (Gilmour ct al..
1998V
In addition to affecting Hg methylation rates, seasonal changes in hydrology can alter
fluxes of Hg in watersheds. In the Marcell Experimental Forest, MN, water flow during
snowmelt over 2 weeks in spring is 30 or 41% of total annual discharge from two wetland
catchments (Mitchell et al.. 2008c'). and 26 or 39% of total annual Hg export from the
catchment occurred during that time. The majority of MeHg within the catchments (79 or
95% of total catchment MeHg) was in the wetlands and exported in high amounts during
snowmelt, about 22-23% of annual MeHg export from the catchments (Mitchell et al..
2008c). In the Adirondacks, there were increases in stream MeHg concentrations and
%MeHg in summer months, when stream discharge was low and water residence time
was long in the wetlands where methylators were active (Gerson and Driscoll. 2016;
Selvendiran et al.. 2008a). High flow events can be important for transporting MeHg
through ecosystems but do not necessarily cause methylation of mercury in aquatic
ecosystems.
Climate Modification of S Effects on Ecosystems
Climate factors that will modify ecosystem response to S include changes in temperature
and intensification of hydrologic cycling (Paraniape et al.. 2017). Recent evidence from
experiments in lakes and bogs allows evaluation of how climate change will affect the
relationship between SOx deposition and MeHg production. Climate change will affect
aquatic and wetland ecosystems via warming temperatures and increasing fluctuations in
water tables due to changes in the timing and magnitude of precipitation (IPCC. 2013).
Generally, microbial activity increases with temperature. It is biologically plausible that
warming could increase MeHg via stimulation of both the decomposer guilds of microbes
producing DOC and of SRPs actively methylating Hg; however, it is also plausible that
warming will increase the activity of microbes that demethylate MeHg. IPCC (2013)
projected that heavy precipitation events will increase in number and strength due to
climate change, which can lead to rapid changes in water table level in wetlands and
lakes. These changes in water level can move pulses of labile carbon and sulfate into the
anoxic zones of the water, sediment, and periphyton where SRPs are active, and flooding
12-45
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
or rewetting of previously dry zones can create new environments favorable for SRPs.
The following studies suggest that drought and warming increase S cycling and Hg
cycling. Drought, but not warming, increases MeHg production in lakes and wetlands.
As reviewed in the 2008 ISA, recovery in Little Rock Lake, WI, has been monitored
because the lake was experimentally acidified for 5 years (1984-1989). Drought affected
sulfate and MeHg concentrations in the lake, which is rain-fed. During the drought years
of 1998-2006, MeHg increased 0.004 ng/L/yr, although high variation among samples
made this trend marginally significant \p = 0.06 ("Watras and Morrison. 2008)1. MeHg
increased from 0.04 ng/L in 1997 to 0.07 ng/L in 2006, a 75% increase, which is more
than can be attributed to evapoconcentration from the lake's 30% decrease in volume
(Watras and Morrison. 2008). During this time period, there were no significant changes
in total Hg deposition; although between 1988 and 2006, the total Hg concentration in
Little Rock Lake has declined 0.04 ng/L/yr (Watras and Morrison. 2008). During the
drought, between 2000 and 2004, total MeHg mass in the north basin of the lake was
positively correlated with higher Hg and S deposition in precipitation (Figure 12-9).
Wetlands are characterized by fluctuating water tables, and microbial communities in
wetlands are adapted to respond rapidly to changes in substrate and oxygen availability.
This is particularly true for SRPs, which methylate Hg in geographic and temporal
pulses, resulting in hotspots of high MeHg concentrations within wetlands where Hg is
actively methylated (Supplemental Material). Fluctuations in water tables caused by
periods of drought and periods of flooding can create pulses of SRP activity and hotspots
of Hg methylation, raising MeHg concentrations in the environment. The following
studies present evidence that climate change (through drought and warming) will interact
with S addition to change MeHg and Hg dynamics in wetlands.
12-46
DRAFT: Do Not Cite or Quote

-------
in

ro
2000 2001 2002 2003 2004
Year
70 80 90 100110 4
HgT (mg- ha"1)
Atmospheric deposition rate
6 8 10
S04 (kg-ha"1)
ha = hectare; HgT = total mercury deposition kg = kilogram; MeHg = methylmercury; mg = milligram; S042- = sulfate.
Source: adapted from Watras and Morrison (2008).
Figure 12-9 Total methylmercury mass in water at Little Rock Lake, Wl,
annually (a), and in relationship to annual mercury (b) or sulfur (c)
deposition.
1	Drought and rewetting increased MeHg in Bog Lake Fen, and S addition treatments
2	enhanced this effect. A 9-month drought from the summer of 2006 to the spring of 2007
3	resulted in a pulse of reduced S when the water levels rose again, and the sulfate
4	concentrations were 147% higher in the S addition treatment, and 78% higher in the
5	recovery treatment, than in the control wetland (Wasik et al.. 2015). Sulfate pulses
July 2018
12-47
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
resulted in proportional increases in Hg methylation, with peak MeHg fraction 150%
higher in the S addition and 60% higher in the recovery treatment ("Wasik et al.. 2015).
The peat experienced drought conditions in summer 2007 as well, and pulses of sulfate in
the experimental treatments that fall (165% higher in S addition, 14% higher in recovery
wetland than in control wetlands) resulted in higher peak MeHg fractions in fall and the
following spring in both the S addition and recovery treatments. Experimental
manipulation of water levels in the different wetland treatments showed that higher water
levels raised sulfate, total Hg, and MeHg in S addition wetlands, but did not affect these
parameters in control and recovery wetlands (Wasik et al.. 2015). In the recovery
treatment, S and MeHg pulses were higher than in control plots, indicating persistent
effects of S addition several years after addition ceased. However, because
experimentally raising water levels did not alter S or Hg chemistry in control or recovery
wetlands, the recovery wetland was returning to S and Hg processes more similar to the
control plots.
Warming may ameliorate the effect of S upon MeHg production in wetlands, but
experimental evidence suggests that it will also increase sulfate and MeHg export from
peatlands into downstream aquatic ecosystems. Degero Stormyr in Sweden is the site of a
wetland experiment that addresses the effects of S and N deposition as well as warming
due to climate change. Warming altered the S and Hg dynamics under S deposition at
Degero Stormyr. Total Hg (per mass of peat) was 19% lower in 20 kg S/ha/yr + warming
plots than in S only plots, indicating increased mobilization or volatilization of Hg from
the peat. The Hg methylation constant in S + warming was similar to the rate in
unamended control plots, much lower than the rate in S only plots (Akerblom et al..
2013). MeHg concentrations in the S and warming treatment were lower than in any
other treatment, 73% lower in the 20 kg S/ha/yr + warming treatment than in the 20 kg
S/ha/yr treatment, and 54% lower than in the control treatment (Akerblom et al.. 2013).
This last result suggests that the combination of warming and S addition increases MeHg
mobility beyond current ambient rates. Given the decreases in both total Hg and MeHg, it
appears that all forms of Hg were mobilized and either exported or volatilized by the
warming + S addition treatment, despite decreases in Hg methylation rate.
12.3.3.4. pH and Methylation Potential
The 2008 ISA described the state of knowledge current at the time that lower pH waters
had higher MeHg production. This relationship was assumed to be causal. However, the
laboratory body of evidence suggests that if all other factors are held equal, methylation
is highest at near neutral pH and drops with increasing acidity. The apparent discrepancy
July 2018
12-48
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
is due to the fact that SOx deposition causes both elevated sulfate concentrations and
acidification in aquatic ecosystems.
Studies from the early 1980s assumed that acidification, not sulfate concentration, was
the causal factor in observed increases of MeHg in water and increased Hg burdens in
fish in response to SOx deposition. For example, samples from circumneutral lakes at the
ELA in Ontario that were experimentally acidified with H2SO4 had increased methylation
and decreased demethylation, so that methylation:demethylation rates were 2.9 to 4 times
higher at pH 5.1 than at pH 7.1 (Xun et al.. 1987; Ramial et al.. 1985).
Recent research on the role of S reducing bacteria in methylating Hg (Appendix 12.3.2)
suggests reinterpretation of an acid pH-MeHg relationship. Controlled laboratory
experiments showed that it was the added sulfate, not pH, that increased Hg methylation
rates in these lakes. Slurries with sediments from Lake Clara, WI, were made with added
H2SO4 or a molar equivalent amount of sulfate as Na2S04. Acidification from neutral
slurries to pH 4.5 with H2SO4 decreased methylation by 65%, while an equivalent
addition of Na2S04 did not alter methylation rates; acidification to pH 3.0 almost entirely
inhibited methylation (Steffan et al.. 1988). This study also measured demethylation over
a range of pH 2.0-8.0. The results showed that demethylation was higher than 2% from
pH 4.4-8, while methylation rose above 2% from pH 6.0-8.0, and both processes slowed
with decreasing pH (Steffan et al.. 1988). The 2008 ISA correctly identified the
correlation between acidic surface water and high rates of MeHg, but the two factors
share a common cause, which is increased surface water sulfate concentrations, rather
than a direct causal relationship. In the Adirondacks, which have historically experienced
high S deposition (see Appendix 16.2). surface water total Hg, MeHg, and %MeHg were
unrelated to pH when all variables were measured across 44 lakes in 2003-2004 (Yuet
al.. 2011). However, there were effects of pH on Hg in biota (see Appendix 12.4). In
regions like the Adirondacks, which historically received high S deposition and have not
fully recovered from this historical acidification, pH remains an important predictor of
Hg in biota.
12.3.3.5. Organic Matter and Methylation Potential
Organic matter is the metabolic substrate of all the microbial methylators, including both
the group of organisms responsive to sulfate and the microbial methylators that depend
on other metabolic pathways. Organic matter in environments where carbon affects
mercury cycling may be quantified as soil organic matter (SOM; as percentage of soil or
in g OM/g sediment); total organic carbon (TOC, g carbon/mL pore or surface water) or
its components; particulate organic carbon (POC, g C/mL water); and dissolved organic
July 2018
12-49
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
carbon (DOC, g C/mL water). The 2008 ISA described some of the complex qualitative
relationships (stimulatory and/or inhibitory) between DOC and rates of Hg methylation.
Organic matter influences the mercury cycle in a number of important ways. Organic
matter affects MeHg production through effects on inorganic Hg ion chemistry and
availability to microbes, and by serving as a metabolic substrate for SRPs. New studies
from lakes, rivers, wetlands, and estuaries have found positive relationships between a
range of lower DOC concentrations and MeHg production or concentrations. However,
the chemical form of DOC or the salinity of the water body can alter DOC and MeHg
relationships.
Certain forms of DOC may serve as a metabolic substrate that increases SRP activity, and
by extension, Hg methylation. In wetlands in the Arbutus Lake Watershed, NY, there
were strong correlations (r2 = 0.58 or 0.60) between Hg and total carbon in the upper
30 cm of peat (Selvendiran et al.. 2008b'). In northern Wisconsin lakes, the quantity of
DOC derived from wetlands is the strongest (>80% variance) predictor of MeHg (Watras
et al.. 2006). In Canadian boreal lakes, there was a positive relationship between
methylation rates and DOC (r2 = 0.62). In Little Rock Lake, WI, DOC and S reduction
had a complex relationship: in the range of 3.6-7.6 mg C/L (reported as
300-600 (imol/L), sulfide (a proxy for S reduction) increased linearly with DOC
concentration, while S reduction did not occur at DOC less than 3.6 mg C/L, and S
reduction did not increase at DOC concentrations above 7.6 mg C/L (Watras et al.. 2006).
In Lake Geneva, Switzerland, mercury methylation rates were an order of magnitude
higher in settling particles than in sediments, and analysis of the particles suggested that
they were derived from algal biomass (Gascon Diez et al.. 2016).
Sediment organic matter may also increase microbial Hg methylation. At another
watershed in the Adirondacks, Sunday Lake, NY, the highest MeHg concentrations were
in samples that were more than 20% organic material (Yu et al.. 2010). In prairie potholes
in Saskatchewan, surface water [MeHg] increased in a linear relationship with the
sediment percentage of organic matter (OM), with a 2.65 ng/L increase in MeHg for
every 10% increase in the percentage of OM (Hoggarth et al.. 2015).
In wetlands, rivers, and lakes of the Mississippi River Delta, LA, total Hg dissolved in
surface water was positively correlated with DOC concentrations, but negatively
correlated with sulfate concentrations. However, this correlation is confounded by
ecosystem salinity, because mean sulfate concentrations in surface water were 16 mg/L in
freshwater wetlands, 63 mg/L in brackish wetlands, and 909 mg/L in marine wetlands
(Hall et al.. 2008). Also, dissolved MeHg in surface water increased linearly with the
hydrophobic organic acid fraction of DOC, with dissolved MeHg increasing 1 ng/L
(0.000001 mg/L) for each increase of 0.048 mg organic acid/L (Hall et al.. 2008). The
12-50
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
hydrophobic organic acid fraction represents a recalcitrant source of aromatic carbon
such as that found in peat, which would account for the high amounts of MeHg in these
wetlands. The hydrophobic organic acid fraction of terrestrially derived C also controls
Hg export (see paragraph below). Another recent study used controlled incubations of an
SRP strain, varying concentrations of Hg, and two well-characterized forms of DOC to
quantify the effects of DOC on methylation. DOC typical of terrestrial sources (high
aromatic fraction, high molecular weights) increased MeHg production 10-40 times
above unamended controls, while aquatic DOC (high aliphatic fraction, low molecular
weights) doubled MeHg production (Graham et al.. 2012). The authors hypothesize that
DOC increases MeHg production by slowing the growth of Hg-S particles, increasing
Hg-S residence time in the water, and enhancing Hg bioavailability (Graham et al..
2012). The stronger effects of terrestrial DOC on MeHg may explain why MeHg
production is particularly high in wetlands, recently flooded reservoirs, and periodically
flooded river plains (Benoit et al.. 2003). In estuarine and marine sediments of the
Chesapeake Bay, MD, organic C in the sediments had a strong positive correlation
(r = 0.96) with sediment MeHg concentrations (Hollweg et al.. 2010). In coastal marshes
of the Chesapeake such as Kirkpatrick Marsh, MD, there was also a statistically
significant positive relationship (r2 = 0.478) between Hg methylation rate and
mineralization of organic C, a metric of cellular metabolism within the sediment
(Mitchell and Gilmour. 2008).
Organic matter in surface water can increase the amount of Hg transported throughout a
watershed, as well as MeHg production at the watershed scale. DOC can form complexes
with Hg ions that facilitate their transport. In lakes and rivers of the St. Louis River, DOC
and dissolved Hg were correlated (r2=0.69), as were aromatic DOC and dissolved Hg
(r2 = 0.70), indicating that aromatic DOC transports Hg through aquatic ecosystems in
this watershed (Jcrcmiason et al.. 2016). In riparian forest soils in Oak Ridge, TN,
flooding resulted in a pulse of both aromatic DOC and MeHg in the soil solution (Poulin
et al.. 2016). In the Arbutus Lake watershed in the Adirondack Mountains, NY, the fluxes
of Hg and DOC in streams were correlated (r2 = 0.80) across seasons, and the highest
fluxes were in June and July (Selvendiran et al.. 2008a). In a study of five watersheds
across the U.S., MeHg concentrations in catchment outflows correlated positively with
particulate organic matter (POC) in Sleepers River, VT (r2 = 0.93), and in Rio Icacos,
Puerto Rico [r2 = 0.83 (Shanlev et al.. 2008)1. In three Northeast watersheds, stream
export of dissolved Hg correlated positively with DOC, with 0.34 ng Hg/L increase for
each 1 mg C/L increase (r2 = 0.87), and hydrophobic organic acid fraction of DOC
explained 91% of variation in dissolved Hg (Dittman et al.. 2010). Storm events, which
increased water flow through surface soil layers and stream discharger, also increased the
export of organic matter and associated Hg (Dittman et al.. 2010). Thus, the evidence
12-51
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
indicates that the amount and form of organic matter will affect Hg transport in
watersheds.
DOC affects the partitioning of inorganic Hg between sediment and water, and hence the
bioavailability of inorganic Hg to methylating organisms (Hsu-Kim et al.. 2013; Marvin-
Dipasquale et al.. 2009; Benoit et al.. 2003). In Kirkpatrick Marsh on the Chesapeake
Bay, MD, linear relationships between methylation rates and absorbance values indicated
that Hg methylation rates were higher when the DOC was composed of high molecular
weight, aromatic organic compounds (Mitchell and Gilmour. 2008). There were no
correlations between DOC characterization and measures of microbial activity, so the
authors suggested that the form of DOC (size and aromaticity of compounds) controls Hg
bioavailability rather than serving as a metabolic substrate powering MeHg production
(Mitchell and Gilmour. 2008). However, the relationship between DOC and Hg
partitioning is complex [reviewed in Hsu-Kim et al. (2013)1. and in an extensive survey
of the Alabama River basin around Mobile, AL, there was no relationship across
52 sampling sites between DOC and Hg fractions [total, aqueous, particulate (Warneret
al.. 2005)1.
12.3.3.6. Iron and Methylation Potential
The 2008 ISA briefly identified iron (Fe) as a surface water component that can alter the
relationship between SOx deposition and MeHg production (as it also interacts with
sulfide phytotoxicity and internal eutrophication, see Appendix 12.2.3 and
Appendix 12.2.4). Iron can alter the rates of Hg methylation via abiotic and biotic
mechanisms. Iron binds with sulfide produced by SRPs during methylation to precipitate
out of solution (Hellal et al.. 2015). preventing the negative feedback of sulfide on sulfate
reduction. Iron oxides form surface complexes with inorganic Hg that decrease Hg
bioavailability. However, sulfide released by SRPs can react with Fe-Hg complexes to
release Hg into the water column (Hellal et al.. 2015). Iron can increase methylation rates
directly, by stimulating the activity of iron-reducing bacteria capable of Hg methylation,
including Geobacter spp. In laboratory incubations mimicking freshwater sediments, iron
reduction and sulfate reduction co-occurred in time but at different zones in the sediment
(Hellal et al.. 2015). Iron and S reduction also co-occurred in wetlands of the Yolo
Bypass, CA (see Appendix 12.3.4.3).
In Kirkpatrick Marsh on the Chesapeake Bay, MD, there was a positive relationship
(r2 = 0.478) between Hg methylation rate and mineralization of organic C and an equally
strong positive relationship between Hg methylation rate and the pool of Fe2+ [r2 = 0.478
(Mitchell and Gilmour. 2008)1. Although Fe precipitation with Hg-S has been posited as
July 2018
12-52
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
one of the mechanisms by which Hg is buried in sediments, there was no relationship
between iron concentration and partitioning of Hg between water and soil.
12.3.3.7. Nitrate and Methylation Potential
The 2008 ISA did not address the effect of nitrate upon MeHg production. Nitrate ions
are energetically favored over sulfate in redox reactions, so N addition can depress Hg
methylation by shifting the energetic advantage from SRPs to denitrifiers (Dcv ct al..
2015). In Florida Everglades sediment incubations, addition of 1.4 mg N/L as NOs to
incubations decreased MeHg production by 70% (Gilmour et al.. 1998). In Onondaga
Lake, NY, the Syracuse water treatment plant incorporated year-round nitrification in
2004, doubling the nitrate concentrations measured annually in May before the lake
stratified (Todorova et al.. 2009). MeHg concentrations decreased exponentially with
increasing NO3 concentrations (estimated pseudo-r2 = 0.61) in the 2-3 years following
the change in water treatment. This change in N load reduced MeHg accumulation in the
hypolimnion by 50% in 2006, and by 93% in 2007 compared to MeHg measured earlier,
in 1982, 1992, and 2002 (Todorova et al.. 2009). In summer 2011, a calcium nitrate
solution was pumped directly into the hypolminion of Onondaga Lake thrice weekly,
adding 84 MT of N over the course of the season (Matthews et al.. 2013). As a result,
hypolimnion N-NO3 remained above 1.0 mg/L during summer stratification, and
hypolimnion maximum MeHg concentration decreased 94% compared to 2009.
(Matthews et al.. 2013) attributed this to two effects of N: a shift in competitive
advantage from SRP to denitrifying bacteria, and increased sorption of MeHg to lake
sediments.
12.3.3.8. Salinity and Methylation Potential
Mercury methylation occurs in estuarine and marine ecosystems (Lchnhcrr et al.. 2011;
Benoitetal.. 1998). but sulfate concentrations in saline water are generally so high that
direct deposition of SOx is unlikely to alter existing MeHg dynamics. An early study on
coastal marshes in Georgia showed that both sulfate reduction and MeHg production
were at their highest rates in the top 4 cm of the marsh soil (King et al.. 1999). An older
study of Hg dynamics in the Patuxent River Estuary, MD, where sulfate concentrations
range from 150 to 1,200 mg/L, found no correlation between sediment pore water sulfate
concentrations and MeHg fraction in sediments (Benoitetal.. 1998). There was a peak in
surface water MeHg fraction of 25% at the mouth of the river, which authors speculated
was due to mixing by wind of stratified anoxic water from the Chesapeake Bay with the
river water (Benoitetal.. 1998). MeHg sediment fraction across the Patuxent River, MD,
July 2018
12-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
was negatively correlated with sulfide sediment concentrations, presumably because Hg
bound to sulfide is unavailable to methylating bacteria (Benoit et al.. 1998).
12.3.4. Mercury Methylation in Sulfur Addition Field Studies
There are a number of ecosystems in which sulfate is elevated by anthropogenic
activities. Scientists have added sulfate to ecosystems in order to study S and Hg cycling
(see Appendix 12.3.4. IV S is used as an agricultural amendment in peat soils, and so Hg
cycling has been extensively studied in the Everglades Water Conservation Area (see
Appendix 12.3.4.2) and the San Joaquin Delta (see Appendix 12.3.4.3V
12.3.4.1. Ecosystem Scale and Field Mesocosm Studies
The 2008 ISA found that experimental S addition to aquatic and wetland ecosystems
increased MeHg concentrations in surface water (and Hg concentrations in biota, see
Appendix 12.4). Key studies supporting this finding focused on the long-term S
acidification (1984-1989) and recovery (1990-present) experiment at Little Rock Lake,
WI and the relatively recent (initiated in 2001) experimental S addition experiment at
Bog Lake Fen in the Marcell Experimental Forest, MN. This section integrates older
published reports and recent papers from S addition experiments that presented
quantitative relationships between S addition and Hg mobilization, methylation, and
MeHg concentrations in surface water. There is new evidence from the S addition
experiment at Bog Lake Fen and from an S addition by warming experiment in a Swedish
bog that confirms the relationship reported in the 2008 ISA between increased S loading
and environmental increases in MeHg. There is also new evidence that decreases in S
addition allow recovery of MeHg toward control conditions, as well as new evidence that
climate change (drought, warming) will increase S cycling and increase MeHg
concentrations (Table 12-7).
As reported in the 2008 ISA, experimental S addition increases MeHg production in lake
ecosystems, and cessation of S addition results in reductions in MeHg production in
lakes. Little Rock Lake, WI was an S addition experiment that ran from 1985-1991. In
this experiment, the north and south basins of the lake were artificially isolated by a
plastic barrier, and then annual S additions were used to acidify the northern basin, while
the unamended south basin served as a control. In 1993, after S addition had ceased but
sulfate concentrations in the north treatment basin (4.8 mg/L or 50 (j,M) were still twice
that of the reference basin (2.4 mg/L or 25 (j,M), the potential methylation rate was 220%
higher in the treatment basin (Gilmour and Riedel. 1995). In June 1993, when MeHg
July 2018
12-54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
fractions were at their highest, the MeHg fraction of total Hg was 30% higher in the
treatment basin (Gilmour and Riedel. 1995). Mercury and S cycling were monitored in
the lake through 2003, and elevated sulfate concentrations and acidification persisted in
the lake through 1996; the lake was considered recovered from 1998-2003. During
acidified years, 25% more of annual Hg2+ inputs were transformed into water column
MeHg than in recovery years ("Watras et al.. 2006). Net methylation during the summer
was 70% higher in acidified than in recovery years, and summertime MeHg accumulation
was 75% higher in the whole lake in acidified years (Watras et al.. 2006). There was a
strong positive correlation between summer MeHg accumulation (1990-2003) and the
interaction of epilimnion sulfate concentration and epilimnion Hg2+ concentration. The
hypolimnion or the anoxic-oxic boundary were likely the sites of Hg methylation,
because the hypolimnion, which was less than 5% of the lake volume, accumulated more
than 70% of the total MeHg mass of the lake in acidified years (Watras et al.. 2006).
Hypolimnion MeHg was associated with hypolimnion S reduction, as there was a
10.8 ng MeHg/L (0.0000108 mg MeHg/L) increase for each 1 mg/L increase in hydrogen
sulfide (H2S; reported as 1.71 pmol/L increase in MeHg for each 1 (imol/L increase in
H2S; r2 = 0.65). Total MeHg in Little Rock Lake, WI increased 14.4 mg for every 1 kg/ha
increase in sulfate deposition (Watras and Morrison. 2008).
Table 12-7 New studies on mercury methylation in sulfur-amended ecosystems.
Type of Additions or Load
Ecosystem (kg S/ha/yr)
Biological and Chemical Effects Study Site
Study
Species Reference
Peat fen Ambient S
deposition:
3 kg S/ha/yr
S addition: low S
(7 kg S/ha/yr
added for total load
of 10 kg S/ha/yr) or
high S
(17 kg S/ha/yr
added for total load
of 20 kg S/ha/yr)
as Na2SC>4, with 0
or 30 kg N/ha/yr as
NH4NO3
At 30-cm depth, total Hg (per peat
mass) in warming + high S was 19%
lower than in high S plots.
Low S decreased total Hg in peat
methylation hotspots by 25 or 31% on a
per-mass or per-volume basis.
The Hg methylation rate constant
increased by 71% in low S and by 400%
in high S.
Peat MeHg concentrations decreased
by 22% in low S and increased by 74%
in high S.
Degero
Stormyr,
Sweden (64°
¦ 09' N, 20°
22' E),
measured in
2007 and
,2008
Peat
samples
Akerblom
et al.
(2013)
Peat fen
Ambient S
deposition:
3 kg S/ha/yr
S addition: high S
(17 kg S/ha/yr
S addition increased pore water MeHg	Degero
concentrations by 117%. S addition	Stormyr,
increased the range of MeHg	Sweden (64°
concentrations by 4.3x, and the	11' N, 19°
variance by 22.2x the control variance.	33' E),
Pore water
in peat
Bergman et
al. (2012)
July 2018
12-55
DRAFT: Do Not Cite or Quote

-------
Table 12-7 (Continued): New studies on mercury methylation in sulfur-amended
ecosystems.
Type of Additions or Load	Study
Ecosystem (kg S/ha/yr)	Biological and Chemical Effects Study Site Species Reference
added for total load With s additioni MeHg increased with measured in
of 20 kg S/ha/yr) higher groundwater levels, although ^
there was no relationship in ambient
plots.
Total Hg in the top 5 cm of peat
increased by 133% with S addition.
Peatland
(poor fen)
Experimental
treatment, addition
of 32 kg/ha/yr in
three simulated
rainfall events,
2001-2008
Recovery
treatment,
32 kg/ha/yr
2001-2006, no
added S
2006-2008
Deposition = 5.5 kg
S/ha/yr (NADP site
MN16)
Mesocosms
installed in
ambient, added S,
and recovery
zones of marsh,
pulse of 130 mg
Na2SC>4 added
In spring 2005, S addition increased
%MeHg 387% over control wetland. In
fall 2005, S addition increased %MeHg
275%.
Following a 9-mo drought, spring 2007
rewetting released internally stored S,
the pulse of SO42" was 78% higher in
recovery and 147% higher in
experimental than in control wetlands.
Following a 9-mo drought and rewetting
in spring 2007, peak [MeHg] was 75%
higher in recovery and 120% in
experimental wetland. Peak %MeHg
was 60% higher in recovery and 150%
in experimental wetland.
Following summer-long drought and
rewetting in fall 2007, [SO42"] was 14%
higher in recovery and 165% higher in
experimental wetland.
Following summer-long drought, the fall
2007 peak [MeHg] was 102% higher in
recovery and 301% higher in the
experimental wetland. Peak %MeHg
was 150% higher in recovery and 350%
higher in the experimental wetland.
Following summer-long drought, the
spring 2008 peak [MeHg] was 62%
higher in recovery and 133% higher in
the experimental wetland. Peak %MeHg
was 146% higher in recovery and 262%
higher in the experimental wetland.
Experimental water level rise in
mesocosms does not affect chemical
constituents in control and recovery
wetlands, but increased total Hg, SO42",
MeHg, and %MeHg in the experimental
wetland. DOC decreases in the
experimental wetland mesocosm.
S6 peatland
in Marcell
Experimental
Forest, MN
Sphagnum
spp.,
herbaceous
forbs,
ericaceous
shrubs,
spruce, and
tamarack
Wasik et al.
(2015)
July 2018
12-56
DRAFT: Do Not Cite or Quote

-------
Table 12-7 (Continued): New studies on mercury methylation in sulfur-amended
ecosystems.
Type of Additions or Load
Ecosystem (kg S/ha/yr)
Biological and Chemical Effects Study Site
Study
Species Reference
Bog	Ambient S load:
2.07 kg S/ha/yr as
quantified by
NADP
S addition:
8.28 kg S/ha/yr
(low), or
20.7 kg S/ha/yr
(high) as SO42"
solution
S addition increased pore water MeHg
concentrations 219-246%.
High S increased MeHg fraction
(MeHg/total Hg) by 117%.
Labile organic C and S additions
increased methylation rates (pg
MeHg/L/day) by 163% and rates of
increasing MeHg fraction (%MeHg/day)
by 152% over rates for S-only addition.
Bog Lake Peat pore Mitchell et
Fen, Marcell water	al. (2008a)
Experimental
Forest, MN
Bog	Ambient S
deposition:
5.5 kg/ha/yr (mid
2000s) as
quantified by
NADP site MN16
S addition:
32 kg S/ha/yr
dissolved in pond
water and
delivered by
sprinklers (mimics
4x the 1990s
deposition rate)
In pore water, MeHg (ng/L) was
increased by 8.8-17.9x the control
levels in 2006 and 3.4-11.7* the control
levels in 2008. MeHg fraction
(MeHg/total Hg) was increased
6.1-13.4x in 2006 and 3.9-11.6x in
2008.
In solid peat, MeHg concentrations and
MeHg fraction were 4-9* higher than in
controls or5-6x higher when
accounting for annual variability.
3 yr after SO42" treatment cessation
(recovery treatment), pore water MeHg
declined 32%, peat MeHg declined
(ng/L) by 62%, and peat MeHg fraction
(%) declined 76%.
S6 peatland, Pore water, Wasik et al.
bog section,
Marcell
Experimental
Forest, MN
peat, and
Culex spp.
(mosquito)
larvae
(2012)
Lakes: rain-
fed LRL,
DL drains
stream fed
by wetland
LRL: [SO42"] =
2.5 mg/L, DL:
[SO42"] = 2.8 mg/L,
stream feeding DL
[SO42"] = 1.6 mg/L
S deposition not
reported
In LRL during drought years, lake SO42"
concentrations increased 0.18 mg/L/yr,
as MeHg concentrations increased
0.004 ng/L/yr.
In stratified basin of LRL, total MeHg
increased with increasing S deposition:
mg MeHg = 14.368 * (kg
S/ha) - 18.494.
In stream draining wetland, SO42"
decreased 80% as MeHg increased
69% over the course of spring and
summer.
Little Rock
Lake (LRL)
and Devils
Lake (DL),
Wl
Lake water
Watras and
Morrison
(2008)
C = carbon; cm = centimeter; DL = Devils Lake; Ha = hectare; Hg = mercury; kg = kilogram; L = liter; LRL = Little Rock Lake;
MeHg = methylmercury; mg = milligram Na2S04 = sodium sulfate; NADP = National Atmospheric Deposition Program;
ng = nanogram; NH4NO3 = ammonium nitrate; S = sulfur; S042" = sulfate; yr = year.
DL = Devils Lake; ha = hectare; kg = kilogram; L = liter; LRL = Little Rock Lake; LTER = Long Term Ecological Research;
MeHg = methylmercury; mg = milligram; ng = nanogram; S = sulfur; S042" = sulfate; yr = year.
Experimental S addition substantially increases MeHg in wetlands. The ELA in Ontario
has been the subject of considerable research in Hg cycling (see Supplemental Material).
July 2018
12-57
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
An early experiment added pulses of sulfate to peat inside collars installed in a poor fen,
and monitored changes in MeHg concentrations over 4 days. In one experiment, addition
of 14 kg sulfate/ha increased MeHg pore water concentrations (ng/L) at the water table
surface 30% above the highest measured MeHg concentrations in the control plot, with a
100% increase in MeHg above control plot concentrations at a 10-cm depth in the water
table (Branfircun ct al.. 1999). Following a second 14 kg sulfate/ha addition, MeHg at
5-cm depth in the peat water table increased to 130% above maximum MeHg in the
control plot. Four days after the first S addition, MeHg at the 5-cm depth remained
elevated at 250% above maximum measured MeHg in control plots (Branfircun et al..
1999). In an experiment the following year, one sulfate addition equivalent to
2.8 kg sulfate/ha was added. After 24 hours, this S addition had increased MeHg pore
water concentrations at the water table surface 100% above maximum measured MeHg in
the control plot, and at the 10-cm depth below the water table, MeHg increased 20%.
This MeHg increase did not persist to 48 hours after addition (Branfireun et al.. 1999).
There is new evidence that addition of labile DOC typical of leachate from a forested
watershed enhanced the stimulatory effect of S addition upon MeHg concentrations in
wetlands. At the Marcell Experimental Forest, MN, field mesocosms were initiated in
Bog Lake Fen to test the effects of S and carbon addition on Hg cycling in an
ombrotrophic peat bog. Mesocosms received experimentally added loads of
8.3 kg S/ha/yr or 20.7 kg S/ha/yr, alone or in combination with additions of organic
carbon in different forms (Mitchell et al.. 2008a'). S addition increased peat pore water
concentrations of MeHg 219-246% above MeHg levels in unamended control plots, with
no significant difference between net methylation under 8.3 versus 20.7 kg S/ha addition
(Mitchell et al.. 2008a'). Adding 20.7 kg S/ha/yr to mesocosms increased the fraction of
the total Hg pool that took the form of MeHg, with MeHg 117% higher than in control
plots (Mitchell et al.. 2008a'). The combined addition of sulfate and organic carbon (as
glucose, acetate, or leachate from coniferous leaf litter) increased methylation rates
(reported in pg MeHg/L/day) by 163% and increased the fraction of MeHg (%MeHg) by
152% over rates measured in mesocosms that received only S addition (Mitchell et al..
2008a'). This study showed that S addition will increase MeHg in surface water in bogs,
particularly when combined with labile C inputs. These results coincide with findings
presented elsewhere in the ISA that DOC concentrations are increasing in surface waters
(Appendix 7) as terrestrial ecosystems recover from acidification.
A decade of research at Bog Lake Fen in Minnesota shows that additional S increases
MeHg production in this wetland, with consequences for water quality and also for
resident organisms (see Appendix 12.4). Bog Lake Fen is the site of a long-term S
addition experiment in which entire sections of the Sphagnum peat wetland have received
additional S loading in simulated wet deposition. As reported in the 2008 ISA, S addition
12-58
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
began in 2001 at a rate of 32 kg S/ha/yr, and increased the levels of MeHg in the bog by
40-190% in the first summer and increased the MeHg export from the bog by 144% in
2002 (Jcrcmiason et al.. 2006). More recent sampling has confirmed the relationship
between S addition and elevated MeHg production and concentrations in the wetland. In
peat pore water sampled 5 cm below the water table for 2 weeks after S addition events, S
addition increased MeHg concentrations by 880-1,790% in 2006 (0.2-0.3 ng MeHg/L in
control, 2.9-4.2 ng/L in S addition plots), and increased MeHg concentrations
340-1,170% in 2008 ("Wasik et al.. 2012). Pore water MeHg fraction (percentage of total
Hg in MeHg form) increased 610-1,340% over control levels in 2006 and 390-1,160%
over control levels in 2008 (Wasik etal.. 2012). Researchers also sampled surface peat
and found that S addition altered MeHg concentrations in peat. MeHg concentrations
(ng/L) and MeHg fraction (percentage of total Hg as MeHg) were 4-9 times higher in S
addition plots than in control plots, although there was no effect of S addition on total Hg
in plots (Wasik et al.. 2012). The elevated MeHg in water and peat had consequences for
the food chain (see Appendix 12.4).
There is also evidence of a link between S addition and MeHg production from a
European wetland experiment. Degero Stormyr in Sweden is the site of an experiment
that addresses the effects of S and N deposition as well as warming due to climate
change. Plots received S loads of 10 kg S/ha/yr or 20 kg S/ha/yr, and a subset of plots
was enclosed by greenhouses to raise air temperature by about 4°C and peat temperature
(at the 20-cm depth) by about 2°C. Four years after the 1995 initiation of the experiment,
Bergman et al. (2012) measured MeHg production in the control and 20-kg-S/ha/yr plots.
S addition increased mean growing season peat pore water MeHg concentrations 117%
over MeHg concentrations in control plots (Bergman et al.. 2012). There was higher
variation in MeHg in S addition plots (20 kg S/ha/yr), as the range of measured
concentrations was 4.3 times the range of measured MeHg in control plots (Bergman et
al.. 2012). The experiment was resampled 12 years after initiation, and the Hg
methylation constant was 71% higher in methylation hotspots (a hotspot was
experimentally defined as the point in each sampled peat profile with the highest Hg
methylation rate constant) under addition of 10 kg S/ha/yr compared to control plots,
while MeHg concentrations were 22% lower than in control hotspots (Akerblom et al..
2013). These seemingly contradictory results (higher production, lower concentration)
may indicate increased mobilization of MeHg from methylation hotspots (see below) or
increased microbial demethylation rates, which were not tested. Under higher S addition
of 20 kg S/ha/yr, the Hg methylation rate constant was 400% higher than in the control
plots, while the MeHg concentrations were 74% higher than in control plots (Akerblom et
al.. 2013). Together, these results indicate that in wetlands an S addition of 10 or
20 kg/ha/yr increases microbial MeHg production and that an S addition of 20 kg/ha/yr
increases MeHg concentrations.
12-59
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
S addition also increases the mobilization and transport of Hg and MeHg from peat
methylation hotspots into the water column, increasing bioavailability. After 4 years of S
addition at Degero Stormyr in Sweden, plots that received 20 kg S/ha/yr had 133% higher
total Hg levels in the top 5 cm of peat than control plots (Bergman et al.. 2012). After
12 years, sampling of Hg methylation hotspots in plots with 10-kg-S/ha/yr addition found
total Hg was 25% lower per peat mass or 31% per peat volume than in control plots
(Akerblom et al.. 2013). These results indicate that S addition mobilizes Hg stored in
deeper parts of the peat profile, resulting in Hg transport to more accessible parts of the
peat profile.
Recent work from the experiment at Bog Lake Fen also provides evidence that reduction
of experimental S addition to wetlands results in reductions of MeHg concentrations in
wetlands. In 2006, after 5 years of elevated S levels, a recovery treatment was initiated in
the S6 bog experiment, where S addition was halted in recovery plots. Following the
cessation of S addition, pore water MeHg (ng/L) declined by 32% between 2006 and
2008, and peat MeHg concentration and peat MeHg fraction declined by 62 and 76%,
respectively ("Wasik et al.. 2012).
12.3.4.2. Agricultural Systems: The Everglades Water Conservation Area, Florida
There are multiple new studies that relate Hg concentrations in water, periphyton, soil, or
fish to sulfate concentrations in the Everglades Water Conservation Area (see
Table 12-8). In the Water Conservation Areas outside Everglades National Park, FL
(Class I area), sulfate concentrations were as high as 48 mg/L near the EAA, whereas in
pristine areas distant from the EAA, sulfate concentrations were less than 0.5 mg/L.
Previous studies noted that Hg methylation was highest in areas of the wetland where
sulfate concentrations ranged from 0.7 to 1.9 mg sulfate/L (Bates et al.. 2002; Gilmour et
al.. 1998). More recently, a study by Corrales et al. (2011). quantified the S budget and
determined that atmospheric deposition of S is 15 kg S/ha/yr, 4% of the total S load in the
WCAs. MeHg concentrations in surface water in the Florida Everglades increase when
sulfate concentrations are above 1 mg/L, which Corrales et al. (2011) recommended as a
target level for sulfate reduction efforts (Figure 12-10).
Table 12-8 New studies on mercury (Hg) and sulfur (S) cycling in the Everglades
Water Conservation Areas (WCA).
July 2018
12-60
DRAFT: Do Not Cite or Quote

-------
Table 12-8 (Continued): New studies on mercury (Hg) and sulfur (S) cycling in the
Everglades Water Conservation Areas (WCA).
Additions	Biological
Type of or Load	and
Ecosys (kg S/ha/y	Chemical Study Study
tem r)	Effects Site Species Reference
Freshw
ater
marsh
W3 pore
water
[S042"]
<4 |jM
U3 pore
water
[S042"]
39 |jM
F4 pore
water
[S042"]
74 |jM
Deposition
= not
measured
I is
I is
SRB
abundance
was 6.9x
higher in U3
than W3,
16.8x higher
in F4.
In W3,
abundance of
SRB and
methanogens
are not
significantly
different. In
U3, SRB
abundance is
80% higher
than
methanogen
abundance. In
F4, SRB
abundance is
60% higher
than
methanogens.
mcrA copy
number
correlates
positively with
mRNA and
methane
production
rates.
Acetotrophic
methanogens
are dominant
at W3, while
at high S U3
and F4 sites,
hydrogenotrop
hie
methanogens
are dominant.
F4, U3,
and W3
sites in
Water
Conserv
ation
Area,
Everglad
es, FL
Methanog
ens as
quantified
by mcrA
copies
Sulfur
reducing
bacteria
as
quantified
by dsrB
copies
Bae etal. (2015)
July 2018
12-61
DRAFT: Do Not Cite or Quote

-------
Table 12-8 (Continued): New studies on mercury (Hg) and sulfur (S) cycling in the
Everglades Water Conservation Areas (WCA).

Additions
Biological


Type of
or Load
and


Ecosys
(kg S/ha/y
Chemical
Study
Study
tern
r)
Effects
Site
Species Reference
Wetland
Total S
MeHg surface
Everglad
Surface Corrales et al. (2011)

load is
water
es
and

110,303
concentration
agricultu
groundwa

mtons/yr,
rises above
ral area
ter

atm dep is
suggested
(EAA),


4% of
threshold level
FL


total, or
of 1 mg/L



15 kg
sulfate in



S/ha/yr as
groundwater.



measured




by US




EPA




CASNET




(FL11 and




FL99)



Constru
Deposition
Fish Hg
Western
Gambusi Fena et al. (2014)
cted
not
(Y= log mg
Palm
a
wetland
reported
Hg/kg G.
Beach
holbrooki
s
S load as
holbrooki) was
County,
(eastern

[S042"] in
negatively
FL
mosquitof

wetland
correlated

ish) and

inflow:
with water

surface

39-110 m
sulfate

water

g/L
(X = log mg



S042"/L)
concentration
s in three
treatment


wetlands:
Y	= -2.09X +
2.60;
Y	= -3.17X
+3.80;
Y= -1.09X
-0.01.
July 2018
12-62
DRAFT: Do Not Cite or Quote

-------
Table 12-8 (Continued): New studies on mercury (Hg) and sulfur (S) cycling in the
Everglades Water Conservation Areas (WCA).
Additions	Biological
Type of or Load	and
Ecosys (kg S/ha/y	Chemical Study Study
tem r)	Effects Site Species Reference
Wetland Deposition Fish Hg (mg Everglad Gambusi Gabriel et al. (2014)
s and
canals
not
reported
S load as
[SO42"] in
surface
water:
0-60 mg/L
Hg/kg or mg
Hg/kg/mm for
Lepomis)
increased
over 0-1 mg/L
SO42",
reached peak
levels at
1-12 mg/L
SO42", and
decreased
over
12-25 mg/L
SO42", to
remain flat at
SO42"
concentration
s 25-60 mg/L.
es
protectio
n area,
FL
a spp.
(mosquito
fish),
Lepomis
spp.
(sunfish),
and
Micropter
us
salmoide
s
(largemo
uth bass)
Constru
cted
wetland
s
Deposition
not
reported
S load in
surface
water as
[S042"]:
39.3-110.
1 mg/L
(mean
59.5 mg/L)
Between
2000-2004,
total Hg and
MeHg wetland
export were
negatively
correlated to
inflow [SO42"].
tHg: (log ng
tHg/L) =
-1.22*(log mg
S0427L) + 2.4
0; MeHg: (log
ng
MeHg/L) = -1.
91*(log
[S042"])
+3.02.
Between
2000-2011,
sulfate was
negatively
correlated
with tHg and
MeHg as one
variable in
multivariate
models.
Western
Palm
Beach
County,
FL
Surface Zheng et al.
water (2013)http://hero.ero.gov/index.cfm?action=search.vie
w&reference id=2044818 - ENREF 2954
atm = atmospheric; CASTNET = Clean Air Status and Trends Network; dep = deposition; EEA = Everglades agricultural area;
ha = hectare; Hg = mercury; kg = kilogram; L = liter; MeHg = methylmercury; mg = milligram; ng = nanogram; S = sulfur;
S042" = sulfate; THg = total mercury; yr = year.
July 2018	12-63	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
6
cn
c
o
(C
at
o
c
o
o
3
o
a_
o
E
4-1
3
2
Threshold

~ ~
a-	4	*	~
0	** I
^ 1	* * *1^ ~~ * #~ «*
1	0 Jfffe
2	4	6	8
Sulfate concentration (mg L-i)
10
L = liter; mg = milligram; ng = nanogram.
Source: adapted from Corrales et al. (2011).
Figure 12-10 The relationship between surface water sulfate and
methylmercury concentrations in the Florida Everglades.
Orem et al. (2011) suggested that at high surface water sulfate (>20 mg/L) concentrations
in the Florida Everglades, sulfide accumulates and inhibits methylation. In treatment
wetlands that collect stormwater runoff from the EAA, the concentrations of sulfate are
so high that sulfide inhibition of MeHg may occur at the higher end of the sulfate range.
Zheng et al. (2013) reported on the water quality parameters, total Hg, and MeHg in the
inflow and outflow waters of a constructed treatment wetland in the Everglades WCAs.
The wetland was a net source of MeHg to surface water between 2000 and 2004, and
during this time, total Hg and MeHg concentrations in water exiting the wetland were
negatively correlated with sulfate concentrations of water flowing into the wetland
(Zheng et al.. 2013). The range of sulfate concentrations of inflow was 39-110 mg/L
sulfate, with mean sulfate of 59.5 mg/L. Zheng et al. (2013) also constructed multivariate
models for total Hg and MeHg, and even with other factors such as DOC, pH, and
chloride in the model, inflow sulfate remained a strong and negative correlate of outflow
July 2018
12-64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Hg concentrations. The positive relationship between sulfate and MeHg in the Florida
Everglades is significant only at low sulfate concentrations (in this study,
39 mg S0427L).
A gradient of nutrient eutrophication (N, P, and S) from agricultural runoff runs from
northeast to southwest across the WCA, with sulfate concentrations around 4.8 mg/L
(reported as 50 |iM) in southern, less disturbed sites (Gilmour et al.. 1998). Sediment Hg
increased from north to south by a factor of 3-4, and MeHg sediment concentrations
increased across the same gradient by a factor of 25 (Gilmour et al.. 1998). A more recent
study in the Everglades suggested that mercury hotspots (elevated Hg in mosquitofish)
were correlated with relatively oligotrophic conditions (low soil total phosphorus, TP)
and native sawgrass sloughs, while fish Hg was lower in sites with higher soil TP with
invasive cattail stands (Julian et al.. 2016a). Surface water sulfate concentrations
measured at the same sites ranged from 0.2-35 mg SC>427L in the Hg hotspots and
6.5-70 mg SC>427L at the nonhotspots (Julian et al.. 2016b). Surface water sulfate and
pore water sulfate were correlated, and the ratio of pore water sulfide:pore water sulfate
correlated with pore water DOC (Julian et al.. 2016b). suggesting that dissimilatory
sulfate reduction (and by extension, a portion of mercury methylation) is controlled by
DOC as well as sulfate in these marshes. In addition to the studies described in this
section, studies in the Everglades WCA have focused on the microbial mechanisms of
mercury methylation (see Appendix 12.3.2) and Hg loads in wildlife (see
Appendix 12.4).
12.3.4.3. Agricultural Systems: Rice in the San Joaquin Delta, California
There is evidence from the California Central Valley that microbial Hg methylation
occurs in managed agricultural wetlands that produce rice, which suggests a route of
MeHg exposure to humans and wildlife. California (particularly the Central Valley where
these studies were conducted) is the second largest producer of rice in the U.S., as well as
an important stop along a major bird migratory route. The USGS coordinated an intensive
research effort on Hg cycling and methylation of Hg in wetlands of the Central Valley of
California, focusing on the Yolo Bypass Wildlife Area, located in the northwestern
region of the San Francisco Bay delta where there are permanent wetlands dominated by
Typha spp. (cattails) and Scirpus spp. (tule) and also agricultural wetlands in which
Oryza sativa (white rice) and Zizania palustris (wild rice) are cultivated. Rice fields go
through annual drying and rewetting cycles. They are flooded in the summer, drained for
fall harvesting, and flooded in the winter to speed decomposition of rice straw and
organic residue.
July 2018
12-65
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
In the Yolo Bypass, CA, permanent and agricultural wetlands were sampled over the
course of a growing season, June 2007-April 2008. About 10% of the 105 water samples
taken over that time period exceeded the U.S. EPA Hg water quality criterion of
50 ng total Hg/L, and total Hg was higher in the center and outflow of agricultural
wetlands than in permanent wetlands (Alpcrs ct al.. 2014).
Across the wetlands, all of the water samples taken for MeHg determination exceeded the
San Joaquin Delta regulatory total maximum daily load (TMDL) limit of
0.06 ng MeHg/L. In fact, MeHg concentrations in all water samples exceeded 1 ng/L,
with maximum measured water MeHg of 37 ng/L at the outlet of wild rice fields during
harvest. Water MeHg concentrations were significantly higher, about twice as high, in the
agricultural fields as in the permanent wetlands (Alpcrs et al.. 2014). and sediment MeHg
concentrations were also higher in the agricultural wetlands (Marvin- D i pasq ualc et al..
2014). The water MeHg fraction ranged from 1-80%, with a median value of 6.4%, and
there were seasonal variations in agricultural but not permanent wetlands. In agricultural
wetlands, the MeHg fraction increased 20 times between June and August in fields where
white or wild rice was grown, and the MeHg fraction increased 5 times in fallow fields
over the same time period (Alpcrs et al.. 2014).
Hg methylation in the permanent and agricultural wetlands of the Yolo Bypass, CA
correlated with S, Fe, and manganese (Mn) reduction. The strongest correlations of
MeHg fraction in the wetlands were with metrics of Mn reduction, not reduction of Fe or
S (Alpcrs et al.. 2014). Sulfate was not a limiting factor of Hg methylation in the
wetlands, although S34S and sulfate:CI in outlet waters indicated that sulfate reduction
occurred in the wetlands, with at least 20% of the sulfate in the water column reduced
(Alpcrs et al.. 2014). In agricultural wetland sediments, mineralization of C by iron
reduction was 2.4 times the mineralization of C by sulfate reduction, indicating that Fe
reduction was more important than S reduction in terms of anaerobic decomposition
(Marvin-Dipasquale et al.. 2014). Furthermore, addition of sulfate fertilizer in rice fields
did not raise measured sulfate reduction rates to the higher levels observed in fallow,
unfertilized fields, indicating that sulfate was not a limiting factor on anaerobic microbial
activity in the agricultural wetlands (Alpcrs et al.. 2014; Marvin-Dipasquale et al.. 2014).
In the Yolo Bypass, CA, seasonal wetland export of MeHg occurs during the winter as
well as the summer. In the summer, newly methylated MeHg is drawn into the root zone
of the sediment by plant transpiration, and in winter, diffusion releases MeHg from the
sediment into the water column (Bachand et al.. 2014). The short-term storage of MeHg
in wetland soils decreases MeHg exports from the wetland, but increases exposure of
wetland fish to MeHg (Windham-Mvcrs et al.. 2014b). Also, MeHg in rice seeds were
correlated with root MeHg (r = 0.90), suggesting that MeHg from the root zone may be
12-66
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
transported through the plant to tissues that are consumed by migrating waterfowl
(Windham-Myers et al.. 2014a'). Research in other systems has shown that MeHg
produced in sediments is transported to rice grains (Paranjapc et al.. 2017). In a study of
rice paddies surrounding industrial emitters of SOx and Hg in Hunan Province, China
(Xu et al.. 2017a). MeHg and %MeHg in rice grains were positively correlated with soil
S, soil MeHg, and atmospheric concentrations of gaseous elemental Hg. MeHg in rice
fields and in rice grains may be an important pathway to human and wildlife Hg exposure
via rice consumption.
Evidence from fish Hg levels in different wetland types suggests that fluctuating water
levels in the agricultural fields, as well as high levels of organic carbon from winter
decomposition of rice straw, raise Hg methylation rates in the rice fields above rates in
the permanent wetlands. Mercury concentrations of penned Gambusia affinis
(mosquitofish) rose 12 times over initial concentration in white rice fields, 5.8 times in
wild rice fields, and 2.9 over initial fish Hg concentration in the permanent wetlands
(Ackcrman and Eagles-Smith. 2010). and these trends were similar to Hg concentration
trends in wild mosquitofish and Menidia audens (Mississippi silversides) caught in the
wetlands. However, trends of Hg concentrations in macroinvertebrates did not follow
trends in fish. There was no difference by wetland type (permanent, white rice, wild rice)
in Hg concentrations of collected water boatmen, Corisella spp., which feed on plant and
algal biomass. Hg concentrations in collected backswimmers, Notonecta spp., which are
predatory, were higher in collections from permanent wetlands than in collections from
cultivated rice wetlands (Ackcrman et al.. 2010).
Drivers of Mercury Methylation under Ambient Conditions
The 2008 ISA presented evidence that elevated sulfate concentrations increased MeHg
concentrations in aquatic and wetland ecosystems. At the time, most studies documenting
this relationship were conducted in lakes and wetlands, and sulfate was often added
experimentally to test the mechanisms of the relationship. Since then, a number of new
studies have shown that increases in ambient sulfate concentrations can produce
increased MeHg concentrations in surface water and sediments in lakes and wetlands (as
well as increased Hg in wildlife, see Appendix 12.4). In addition, there is new evidence
that increased sulfate in rivers, streams, and watersheds increase Hg methylation. Not all
studies include information about the form or relative contribution of S sources in these
ecosystems, but the studies provide evidence that current concentrations of surface water
sulfate contribute to Hg methylation in a broad range of American aquatic and wetland
ecosystems.
12-67
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
The 2008 ISA described watersheds with low pH waters and high wetland cover as
sensitive to Hg methylation, but did not describe MeHg fraction in the environment as an
indicator of Hg methylation. MeHg is bioavailable and bioaccumulates within organisms
due to the strong complexes it forms with thiosulfate and sulfhydryl groups in organic
molecules. The fraction of total Hg that is in the form of MeHg (or %MeHg) is used as an
indicator of bioavailable Hg within an ecosystem. New studies (as well as key older
studies) of MeHg fractions in watersheds, streams, lakes, and wetlands show the link
between S deposition and MeHg fraction in water and sediments, and confirm the
importance of wetlands and nonurban watersheds as sources of MeHg to aquatic
ecosystems.
12.3.5.1. Sulfur in Ambient Conditions
Aquatic and wetland ecosystems which have received high loads of S generally have
higher MeHg fractions. In sampling of Little Rock Lake and four nearby lakes in
Wisconsin, MeHg fractions in lake water ranged from 6-25% MeHg, with no
relationship to pH (Bloom et al.. 1991). In the Adirondacks, which have historically
experienced high S deposition (see case study), surface water total Hg, MeHg, and
%MeHg were unrelated to pH when all variables were measured across 44 lakes in
2003-2004. Surface water MeHg ranged from 0-48% MeHg, with a mean of 6% (Yuet
al.. 201IV
At the time of the 2008 ISA, only one paper had considered the links between measures
of SOx deposition and MeHg concentrations in ecosystems. Since then, two additional
papers have been published documenting the effects of SOx deposition on Hg levels in
water or fish (Table 12-9). The large scale and scope of these studies include
confounding and interacting environmental factors, such as correlated Hg deposition and
variations in wetland cover, the full effects of which researchers are unable to fully
quantify. Despite this limitation, each of these papers show evidence of temporal or
spatial correlations between SOx deposition and MeHg concentrations in fish or water.
July 2018
12-68
DRAFT: Do Not Cite or Quote

-------
Table 12-9 New studies on correlations between sulfur (S) and methylmercury
(MeHg) in ecosystems.
Type of
Ecosystem
Additions or Load
(kg S/ha/yr)
Biological and
Chemical Effects
Study Site
Study
Species
Reference
River and reservoir
S addition or
deposition not
reported
[SO42"] in Rio
Grande: pore water:
(means)
74-230 mg/L; deep
water: (range)
152-342 mg/L
SO42"] in Devils
River: pore water:
(means)
74-230 mg/L; deep
water: (range)
15-168 mg/L
Sediment MeHg
(ng/g) correlated
negatively with pore
water sulfate
concentrations.
RG had 40% higher
THg than DR.
DR sediment MeHg
was 65% higher than
in RG sediment, and
DR sediment %MeHg
was 65% than in RG
sediment.
Rio Grande
(RG) and
Devils River
(DR) branches
of Amistad
International
Reservoir, TX
Sediment,
pore water,
and deep
water;
Micropterus
salmoides
(largemouth
bass)
Becker et al.
(2011)
Prairie potholes
(shallow freshwater
marsh)
Lowland ponds,
average [SO42"]
14 mg/L
Upland ponds,
average [SO42"]
1,490 mg/L
Deposition not
reported
In July, surface water
[MeHg] was 290%
higher in upland
ponds than in lowland
ponds. In August,
surface water [MeHg]
was 233% higher in
upland ponds.
Surface water %MeHg
was 185% higher in
upland ponds, and
sediment %MeHg was
205% higher in upland
ponds.
Surface water [MeHg]
increases with
sediment organic
matter: MeHg
(ng/L) = -2.26 + 0.265
(percentage OM).
St Denis
National
Wildlife Area,
Saskatchewan,
Canada
Four lowland
ponds with
high SO42"
Five upland
ponds with
low SO42"
Hoqqarth et al.
(2015)
Wetlands
Deposition not
reported
Ambient wetland
[SC>42"]<5 mg/L
Mine-impacted
wetlands,
[S042~]>500 mg/L
Peat %MeHg
increased with soil
acid volatile sulfide in
ambient wetland.
In mine-impacted
wetlands, peat
%MeHg decreased
with increasing S
reduction.
Two wetlands,
St Louis River
Watershed,
MN
One wetland is
downstream of
a mine and
receives high
SO42" effluent
Peat
samples
Johnson et al.
(2016b)
July 2018
12-69
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Table 12-9 (Continued): New studies on correlations between sulfur (S) and
methylmercury (MeHg) in ecosystems.
Type of
Ecosystem
Additions or Load
(kg S/ha/yr)
Biological and
Chemical Effects
Study Site
Study
Species
Reference
Bogs
S addition or
deposition not
reported
Peatland pore water
median [SO42"]: 0.1,
0.4, 0.6, 3.0 mg/L
Pore water [SO42"]
(mg/L) was positively
correlated with pore
water MeHg
(Spearman R = 0.237)
and total Hg
(Spearman
R = 0.576).
two
Four
peatlands
in Marcell
Experimental
Forest, MN;
two in
Experimental
Lakes Area,
Ontario
Pore water
in peat
Mitchell et al.
(2008b)
Upland-peatland
interface
S deposition not
reported
Peat pore water
[SO42"] = 0-20 mg/L
MeHg was positively
correlated with sulfate
(Spearman R = 0.39):
ng
MeHg/L = 0.037 * (mg
S0427L) + 0.58.
S2 and S6
peatlands,
Marcell
Experimental
Forest, MN
Peat pore
water
Mitchell et al.
(2009)
Streams
S deposition not
During the growing
Archer Creek,
Stream
Selvendiran et

reported
season, MeHg
Adirondacks,
water
al. (2008a)

Monthly [S042"]
increased with
NY



means in streams:
decreasing sulfate




4-8 mg/L
concentrations: ng




MeHg/L = -0.11 x (mg
S042"/L) + 0.88.



Streams and lakes Not reported
MeHg was negatively	Arbutus Lake
correlated with water	and Sunday
sulfate concentrations:	Lake,
ng	Adirondacks,
MeHg/L = 6.67 x (mg	NY
S042"/L) - 2.40.
Stream and
lake surface
water
Selvendiran et
al. (2009)
DOM = dissolved organic matter; DR = Devils River; g = gram; ha = hectare; kg = kilogram; L = liter; MeHg = methylmercury;
mg = milligram; ng = nanogram; OM = organic matter; RG = Rio Grande; S = sulfur; S042" = sulfate; THg = total mercury;
yr = year.
A study of fish and SOx deposition sediment records at Isle Royale in Lake Superior
indicated a temporal relationship in this remote location between S deposition and Hg
loads in fish. Isle Royale is a Class I area and contains several lakes impacted by S and
Hg deposition. Between 1981 and 2006, S deposition in the region fell 40-60% (from
about 10 kg S/ha/yr in 1981 to about 6 kg S/ha/yr in 2006 atNADP site MN18 in
mainland Minnesota), while Hg deposition remained stable. S deposition on the island
between 1985 and 2006 was in the range of 2.3 to 7 kg S/ha/yr with a trend (p = 0.09)
toward a decline in deposition over that time period (Drevnick et al.. 2007). Fish
collections were conducted in eight Isle Royale lakes in 1995-1996 and again in
2004-2006. In six lakes, 1995 collections of northern pike (Esox lucius) exceeded the
U.S. EPA fish tissue Hg limit (0.3 |ig Hg/g wet weight). When these lakes were
resampled in 2004-2006, northern pike mean Hg levels were below the U.S. EPA limit,
July 2018
12-70
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
which the authors attributed to a decline in SOx deposition despite steady Hg deposition.
This study also used museum collections of fish from Isle Royale (collected in 1905,
1929, 1966) and an analysis of lake sediment cores to test the relationship between
historical S loads and fish Hg burdens. The chromium-reducible sulfur (CRS) fraction in
sediment cores was determined to closely track known patterns of S deposition since
monitoring began in 1985, allowing the inference of S loads to island lakes since the
Industrial Revolution in the late nineteenth century. The historical S load data were
compared to Hg loads in both the historical and recent fish collections from Isle Royale
lakes. The CRS fraction explained 79% of the variation in Hg concentration of northern
pike fish fillets. There were similar correlations between sediment reduced S and Hg
burden in other fish species, but sample sizes for those species were low and correlations
were not statistically significant (Drevnick et al.. 2007). DOC and pH were not directly
tested for their relationship with Hg levels in fish, but the authors note that there are no
significant trends in DOC or pH in these lakes from 1980-2006. This study showed that
Hg fish burdens increased in Isle Royale lakes with increasing SOx deposition in the
twentieth century, and that Hg fish burdens declined in response to reductions in SOx
deposition since the 1990s.
The other two studies that find effects of SOx deposition are summarized in the following
sections. A 12-year study of MeHg and fish Hg in Voyageurs National Park found that
declines in SOx deposition have resulted in MeHg declines only in lakes where DOC has
remained stable (see Appendix 12.3.5.3). A study that compiled data from reservoir fish
sampling in Texas found correlations between geographic variations in SOx and Hg
deposition and average fish Hg burden over 25 years (see Appendix 12.3.5.2).
Sampling at small scales within peatlands has shown positive relationships between
sulfate concentrations and MeHg concentrations (Table 12-9). In two peatlands at the
Marcell Experimental Forest, MN, researchers monitored water quality at the
upland-peatland interface, hypothesized to be an important zone for methylation due to
labile carbon inputs in runoff from terrestrial communities to the peatland microbial
community. Sulfate peat pore water concentrations ranged from 0-20 mg sulfate/L, and
MeHg concentrations increased 0.037 ng/L for each 1 mg/L increase in sulfate (Mitchell
et al.. 2009). The relationship between sulfate and MeHg is also significant at a larger
spatial scale in the same region. Across four peatlands at the temperate-boreal boundary
in Minnesota and Ontario, MeHg was positively correlated with sulfate in peat pore water
(Spearman's R = 0.237) as well as with pH during the 2005 growing season (Mitchell et
al.. 2008b). The range of median sulfate levels in these peat bogs during the course of the
study was 0.01-3 mg/L, while the range of MeHg concentrations was 0.35-0.62 ng
MeHg/L (Mitchell et al.. 2008b).
12-71
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
The South River in Waynesboro, VA, was contaminated by historical (1929-1950)
textile industrial use and release of Hg. In 2008, sediments were collected within the
channel at 10 study sites at increasing distances downstream of the historic Hg source.
MeHg concentrations in the sediment increased linearly with increasing sediment pore
water sulfate concentrations, with a 0.06 ng increase in MeHg per gram of sediment for
every 1 mg/L increase in sulfate pore water concentration, with mean pore water sulfate
concentrations ranging from 2-16 mg/L (Yu et al.. 2012).
Chemical constituents measured at watershed outflows represent the cumulative
microbial and abiotic chemical transformations within the watershed. A watershed in
which SRPs are present will therefore have higher MeHg and lower sulfate
concentrations in its outflow than an adjacent watershed without active SRPs that
experiences the same S and Hg deposition loads. Many of the following studies document
inverse relationships between sulfate concentrations and MeHg concentrations in
outflows of trunk streams (those that drain an entire watershed) where sulfate and Hg
concentrations reflect upstream sulfate stimulation of SRPs and Hg methylation. In a
tributary stream draining a wetland in Wisconsin, there were seasonal trends in sulfate
and MeHg concentrations indicative of wetland Hg methylation. Sulfate in the stream
declined 80% between April and June 2003, indicating S immobilization or reduction in
the wetland during the early summer, and MeHg increased 69% over the same time
period ("Watras and Morrison. 2008). This increase in MeHg and simultaneous decrease
in sulfate in stream water concentrations suggests upstream activity of SRB in Hg
methylation.
In streams draining wetlands in the Archer Creek watershed in the Adirondack
Mountains, NY, water sulfate concentrations were lower than in streams upstream of the
wetlands (Sclvcndiran et al.. 2008a'). Wetland microbial communities reduced sulfate
while methylating Hg, which would account for the elevated MeHg concentrations in
wetland-draining streams compared to upland streams. There was a negative correlation
between sulfate concentrations and MeHg concentrations in streams draining the
wetlands where Hg methylation occurred, with a 0.11 ng/L increase in MeHg
concentrations in the stream for each mg/L sulfate removed from surface water
(Sclvcndiran et al.. 2008a'). Selvendiran et al. (2009) measured and calculated Hg budgets
for the adjacent watersheds of Arbutus Lake and Sunday Lake in the Adirondacks, NY,
measuring inlet and outlet streams at both water bodies as well as lake surface water at
Arbutus. Across both watersheds, MeHg concentrations were inversely correlated with
sulfate concentrations. This pattern was driven by the concentrations measured in the
streams in Sunday Lake, which drains a watershed that is approximately 20% wetlands
(Selvendiran et al.. 2009). indicating that SRB in upstream wetlands may have been
reducing sulfate and methylating Hg to impact the measured stream surface water.
12-72
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Arbutus Lake drains a watershed of which wetlands comprise only 4% of surface area,
which may explain the relatively higher sulfate concentrations and the lower mean and
variability of MeHg (Sclvcndiran et al.. 2009). In watersheds where S and Hg deposition
rates do not differ spatially, negative correlations between sulfate concentrations and
MeHg show the link between sulfate reduction and Hg methylation.
Highly elevated sulfate concentrations, as can result from agricultural runoff, can depress
SRP activity and Hg methylation, as sulfide produced by SRPs accumulates and
downregulates SRP activity. In a sampling study of the Amistad International Reservoir
on the Texas-Mexico border, Hg pools in sediment, water, and largemouth bass (for more
on this endpoint, see Appendix 12.4) were quantified and compared between the Rio
Grande and Devils River branches of the reservoir. The Rio Grande branch had sulfate
concentrations an order of magnitude higher (site with highest sulfate concentration:
230 mg/L sediment pore water), as well as 40% higher total Hg pools in sediments, than
did the Devils River branch [site with highest sulfate concentrations: 23 mg/L (Becker et
al.. 2011)1. However, total Hg and sulfate were negatively correlated with MeHg, and
SRB (researchers used techniques to detect bacterial sulfur reducers) were present at all
sites. Hg methylation in this system was not limited by Hg supply or biota, and was not
stimulated by the Rio Grande's high sulfate levels, originating from agricultural land use
in the watershed. The Devils River branch had 65% higher MeHg concentrations in
sediments, and a 230% higher proportion of total sediment Hg in the form of MeHg
(Becker et al.. 2011). Becker etal. (2011) posited that the higher MeHg in Devils River
sediments was due to higher DOC in this area of the reservoir and to sulfate
concentrations (in sediment pore water, 2.5-23 mg/L) being optimal for the metabolism
of SRB.
Two wetlands in the St Louis River Watershed, MN, were the subject of a study of Hg
and S because one wetland regularly receives high-sulfate mine effluent (Johnson et al..
2016b). Peat %MeHg increased with S reduction (quantified as pore water sulfide or soil
acid volatile sulfide) in the wetland that did not receive mine effluent, and in which
surface water sulfate concentrations averaged less than 5 mg/L. In wetlands that received
sulfate mine effluent of >500 mg SO42 /L, peat %MeHg decreased with increasing S
reduction. Together, these results support the model of methylation first proposed for the
Everglades [see Figure 12-7. and Glimour (2011)1. in which there is a net methylation
optimum at low moderate sulfate concentrations (1—10s of mg SCV/L) and a
suppression of net methylation at high sulfate (100s of mg SCV/L) and sulfide
(0.6-0.7 mg S/L in pore water) concentrations (Johnson et al.. 2016b).
In contrast, in the St. Denis National Wildlife Area in Saskatchewan, Canada, prairie
potholes vary in sulfate concentrations in relation to geology and position in the
12-73
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
landscape. In a study of water MeHg, five upland ponds had average sulfate
concentrations of 1,490 mg/L (S sources are geologic), and four lowland ponds had
average sulfate concentrations of 14 mg/L (Hoggarth ct al.. 2015). MeHg fraction was
185% higher in surface water and 205% higher in sediments in the upland ponds than in
the lowland ponds, which had lower sulfate concentrations (Hoggarth et al.. 2015).
12.3.5.2. Mercury and Sulfur Interactions in Ambient Conditions
It can be difficult to tease apart the effects of Hg and SOx deposition upon mercury
methylation rates or concentrations of MeHg in biota, because at some scales, Hg and
SOx emissions are correlated (Figure 12-11. Figure 12-12. and Table 12-10). In a
small-scale (15 km, single sampling event) gradient study of Hg, MeHg, and S downwind
of a coal plant and other industrial sources in China (Xu et al.. 2017a). soil S and gaseous
elemental Hg (GEM) were positively correlated, with effects upon cultivated rice. In
surrounding rice paddies, MeHg and %MeHg in rice grains were positively correlated
with soil S, soil MeHg, and atmospheric concentrations of gaseous elemental Hg
(Figure 12-11).
Regionally, Hg and S depositions were correlated in Texas (Drenner et al.. 2011) and
atmospheric reactive Hg and SO2 were correlated in the Ohio River valley (Yatavelli et
al.. 2006). At a national or international spatial scale, different factors govern long-range
transport and deposition of SOx, PM, and Hg, so S deposition and Hg deposition may not
be spatially correlated. However, at a longer temporal scale, Hg and S deposition have
declined in concert over the past decades in the U.S. [U.S. emissions of Hg declined 79%
and SO2 emissions declined 73% between 1990 and 2011 (U.S. EPA. 2017a)l. Temporal
and spatial correlations between Hg and S emissions and deposition make it difficult to
distinguish between Hg and SO42 as the limiting factor in ecosystems receiving high
deposition.
July 2018
12-74
DRAFT: Do Not Cite or Quote

-------
5.0
4.0-
3.0-
SjQ
bO
i
tsfi
~ 2.0
O
o
2
1.0-
0.0
0.0
r=0.56, p<0.05
.-'-'Q °o
r-0.68, p<0.01
O
o
• Rice McHg
O Rice MeHg ratio
-80
60 o
«d
t-H
bX)
40 K

-------
Table 12-10 New studies on deposition of sulfur (S) and mercury (Hg) and their
effect on methylmercury (MeHg).
Type of
Ecosystem
Additions or
Load (kg
S/ha/yr)
Biological and Chemical	Study
Effects	Study Site Species
Reference
Reservoir NADP quantifies Mean Hg fish levels were
S deposition in
2008 at CT:
7.7 kg S0427ha,
TBP: 8.1 kg
S0427ha, ECTP:
8.1 kg S0427ha,
SCTP: 11.9 kg
S0427ha
61-89% higher in the SCP
region (highest S and Hg
deposition) than in other
regions.
145 reservoirs
in four Level 3
ecoregions of
eastern
Texas: CT,
TBP, ECTP,
SCTP
Micropterus
salmoides
(largemouth
bass)
Drenner et al.
(2011)
Lake	S deposition not Between 2004-2015, Hg°
reported	deposition decreased 25%,
while Hg2 deposition remained
constant.
Sulfate concentrations in
Arbutus Lake declined
approximately 40% between
2004-2016. In 2004, sulfate
concentrations ranged
4.3-7.0	mg/L, and in 2016,
sulfate concentrations ranged
2.4-4.3	mg/L.
Between 2004-2015, lake
MeHg and THg concentrations
and fluxes decreased
statistically significantly (mean
changes not reported due to
high variability).
Arbutus Lake
watershed,
Adirondacks,
NY
Water Gerson and Driscoll
samples (2016)
Agricultural S and Hg
wetland deposition not
recorded
MeHg and %MeHg in rice grains Gradient of S Oryza
were positively correlated with
soil S, soil MeHg, and
atmospheric concentrations of
gaseous elemental Hg.
and Hg
deposition
2-15 km
downwind of
Yueyang
Coal-fired
Power Plant,
Hunan
Province,
China
Xu et al. (2017a)
sativa (rice)
grown in
paddies
CT = Cross Timber; ECTP = East Central Texas Plains; ha = hectare; kg = kilogram; MeHg = methylmercury; NADP = National
Atmospheric Deposition Program; S = sulfur; SCTP = South Central Texas Plains; S042" = sulfate; SOx = sulfur oxides;
TBP = Texas Blackland Prairie; yr = year.
1	A study that compiled data from reservoir fish sampling in Texas found correlations
2	between geographic variations in SOx deposition and average fish Hg burden over
3	25 years. Largemouth bass (Micropterus salmoides) were collected between 1985 and
July 2018
12-76
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
2009 with heavy sampling between 2004 and 2008 (45% of the samples) from
145 reservoirs in eastern Texas. The samples were grouped and analyzed by Level 3
ecoregions which received varying amounts of Hg and S deposition, increasing from the
westernmost Cross Timber region where S deposition was 7.7 kg/ha to the easternmost
South Central Plains region where S deposition was 11.9 kg/ha (Premier et al.. 2011).
The South Central Plains region received high deposition of SOx and Hg, and also had
the highest percentage of land area as wetlands (12%) and the lowest percentage of land
area in agriculture of the four regions, all of which are conducive to high methylation
rates in aquatic ecosystems. Mean Hg levels in sampled largemouth bass were 61-89%
higher in the South Central Plains than in other eastern Texas regions (Premier et al..
2011). These results have regional implications because the South Central Plains
ecoregion extends into Arkansas, Louisiana, and Oklahoma. This study showed that
within Texas, the geographic region with the most wetlands and the highest SOx
deposition also had fish with the highest Hg burden.
In the Adirondacks, a temporal study of water MeHg, water sulfate concentrations, and
mercury deposition reported no correlation between stream sulfate and MeHg (as
concentration or as %MeHg) in the summer when SRB are active, in lake data collected
in 2004-2005 (Gcrson and Driscoll. 2016). Over the decade in which these data were
collected, Hg° atmospheric concentration and deposition in the watershed, litter Hg
concentration, and sulfate concentrations in streams all declined in this system; Gerson
and Driscoll (2016) suggested that decreases in surface water MeHg in the watershed
were attributable to decreased Hg deposition, and that surface water sulfate concentration
over the entire decade was so high in the lake that it did not control Hg methylation.
12.3.5.3. Organic Matter and Sulfur Interactions in Ambient Conditions
There is evidence from recent research of interactions between organic matter and sulfur
in controlling mercury methylation (Table 12-11). A 12-year study of MeHg and fish Hg
in Voyageurs National Park found that declines in SOx deposition have resulted in MeHg
declines only in lakes where DOC has remained stable. In lakes in Voyageurs National
Park, MN, a Class I area, researchers measured MeHg in the epilimnion of four lakes
between 2000 and 2012 and compared trends in lake water Hg concentrations with trends
of measured Hg and S deposition (Brigham et al.. 2014). S deposition to the region
decreased 48% between 1998 and 2012 based on regression analysis of data measured at
National Atmospheric Deposition Program (NADP) stations MN16 and MN18, and Hg
deposition decreased 32%. In two lakes, MeHg surface water concentrations decreased
over 12 years of measurement, 50% in Peary Lake and 43% in Ryan Lake (Brigham et
al.. 2014). and fish Hg concentrations followed suit (see Appendix 12.4). In contrast,
July 2018
12-77
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Brown Lake MeHg concentrations increased 85% over the same time period, which the
authors ascribed to increasing inputs from upstream methylating aquatic environments.
An alternative explanation may be in the differences between lakes in dissolved organic
carbon (DOC) concentrations; in Peary and Ryan Lakes, DOC concentrations remained
flat over the decade in which MeHg concentrations decreased, while in Brown Lake,
DOC concentrations increased 30% (Brigham et al.. 2014). Percaflavescens (yellow
perch) were collected between 2000 and 2012 to determine fish Hg levels in comparison
with Hg lake water trends and S deposition. Between 2000 and 2012, P. flavescens Hg
concentrations decreased 37% in Peary Lake and 32% in Ryan Lake, just as S deposition
and lake MeHg declined (Brigham et al.. 2014). In Brown Lake, where DOC increased
30% over the same time period, P. flavescens Hg increased 80% (Brigham et al.. 2014).
In Voyageurs National Park, lakes with stable DOC responded to decreased S deposition
with decreasing MeHg in water and Hg concentrations in fish.
There is evidence from recent research that higher sulfate concentrations correspond to
higher MeHg concentrations and sulfate reduction rates in prairie potholes . Prairie
potholes are shallow wetlands and lakes in depressions created by glaciation, which
depend upon snowmelt and precipitation for their water supply, and tend to be
hydrologically isolated. A survey of prairie pothole wetlands and lakes in Saskatchewan
measured Hg levels in water. Surface water MeHg correlated positively with sulfate
concentrations and negatively with conductivity and specific ultraviolet absorbance
(SUVA; an indicator of aromatic fraction of DOM), such that surface water MeHg
increased 0.03 ng/L for each mg/L increase in sulfate concentrations, when all other
factors remained constant (Hall et al.. 2009a).
In wetlands in the Allequash Watershed, WI, potential methylation rates in streambed
sediments were positively correlated with microbial sulfate reduction (quantified as
sediment acid volatile sulfide, r2 = 0.258), DOC, and reduced iron (Creswell et al.. 2017).
These correlations reflect the activity of methylating SRB, IRB, and methanogens, which
shift in dominance spatially and seasonally across the wetland streams (see
Appendix 12.3.2). Hg concentrations did not correlate with mercury methylation in this
system (Creswell et al.. 2017). The total dissolved S and sulfide fraction did not correlate
with methylation rate, suggesting that sulfate was not a limiting factor, and that DOC was
probably the main driver of methylation in this system.
12-78
DRAFT: Do Not Cite or Quote

-------
Table 12-11 New studies of Interactions between organic matter and sulfur (S) in
Controlling methylmercury (MeHg).

Additions or




Type of
Load (kg
Biological and Chemical

Study

Ecosystem
S/ha/yr)
Effects
Study Site
Species
Reference
Lakes
6.44 kg S/ha in
As SOx deposition decreased
Voyageurs
Surface
Briqham et al.

1998, 3.35 kg
48% in 1998-2012, water MeHg
National Park,
water and
(2014)

S/ha in 2012
decreased 50% in Peary Lake
MN
Perca


as quantified
and 43% in Ryan Lake. In Brown

flavescens


by NADP sites
Lake, water MeHg increased

(yellow


MN16 and
85% in 1998-2012.

perch)


MN18
Between 2000 and 2012, Hg
concentration in P. flavescens
decreased 37% in Peary Lake
and 32% in Ryan Lake, and
increased 80% in Brown Lake.



Wetland
Pore water
Methylation was correlated with
Allequash
Pore water
Creswell et al.

concentrations
acid volatile sulfide (AVS)
Creek, Wl

(2017)

0.0 to 59.2 |jM
(r2 = 0.258, p = 0.001), and




S042"
stepwise with DOC, and
porewater Fe(ll), THg, MeHg and
dissolved Fe (r2 = 0.5096,
p = 0.004).



Reservoir
Deposition not
reported
Permanently
inundated
sediment
porewater:
3.8 mg S042"/L
Seasonally
inundated:
8.2 mg S042"/L
Seasonal inundation increased
sediment MeHg and MeHg
fraction of total Hg in comparison
to permanently inundated
sections:
Seasonal: 2.70 ng MeHg/g
sediment, 0.55% MeHg
Permanent: 0.71 ng MeHg/g soil,
0.18% MeHg
Cottage Grove
Reservoir, OR.
Sediment,
pore water,
and surface
water
Ecklev etal. (2017)
Lakes and
Not reported
Surface water MeHg was
49 lakes and
Lake water Hall et al. (2009a)
wetlands

positively correlated with sulfate
wetlands,



concentrations: ng
prairie pothole



MeHg/L = -2.94 - 0.44
region,



(conductivity) + 0.303(S042") -
Saskatchewan,



0.268 (aromatic fraction of DOM).
Canada

Artificial
Concentrations
Strong correlation noted between
Reservoirs in
Surface Noh et al. (2016)
reservoir
ranged from
sulfate and chlorophyll a
South Korea
water

3-12 mg/L
(r = 0.73), DOC (r = 0.92), and



S042"
conductivity (r= 0.83)


Reservoirs have been identified as sources of MeHg. Cottage Grove Reservoir, OR, was
built for flood control and experiences large, seasonal fluctuations in water levels. Hg and
MeHg dynamics differ between permanently flooded portions of the reservoir and areas
that are inundated during summer months (Ecklev et al.. 2017). When the reservoir was
July 2018
12-79
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
flooded in the spring, sulfate, inorganic Hg, and DOC were released in high
concentrations from previously exposed sediments, creating zones of high mercury
methylation in seasonally inundated sediments. Across nine reservoirs in South Korea,
algal blooms stimulated methylation of mercury. Algal blooms released large pulses of
labile C, as demonstrated by a positive correlation of Chi a and DOC (r = 0.73) (Nohet
al.. 2016). Sulfate concentrations ranged from 3 to 12 mg/L, and also correlated with
DOC (r = 0.92). Concentrations of MeHg and %MeHg were positively correlated with
Chi a, DOC, and sulfate across reservoirs.
12.3.5.4. pH and Sulfur Interactions in Ambient Conditions
Sulfate deposition is not the primary driver of MeHg concentrations in every ecosystem
with elevated Hg concentrations in its biota. In Wallowa-Whitman National Forest, OR
and ID, fish Hg sampled in high-elevation lakes in 2011 correlated positively with basal
area of conifers in the lake catchment and correlated negatively with lake area, surface
water SO42 . and DOC (Eagles-Smith et al.. 2016). This study did not test microbial
methylation rates or community composition, but lakes with low SO42 and low DOC
were acidified, as quantified by pH measurements. Eagles-Smith et al. (2016) interpreted
model results to show that acidification of lakes had negative effects upon fish condition
and resulted in high Hg bioaccumulation, with conifer-forested catchments most at risk.
In Kejimkujik National Park, Nova Scotia, regressions and model selection of Hg in biota
against S34S, S13C, and S15N suggested that lake productivity was a stronger driver of Hg
in biota than sulfur reduction (indicated by enrichment of S34S), which was negatively
correlated with Hg in lake biota (Clavden et al.. 2017). The lakes in this study are
unstratified, shallow and acidic, with S deposition enriched in S34S from nearby marine
waters, all conditions unfavorable to sulfate-reducing bacteria. Clavden et al. (2017)
suggested that microbial methylators not utilizing sulfate are responsible for producing
MeHg in these lakes, and that trophic interactions are responsible for variations in Hg in
biota across lakes.
12.3.5.5. Nutrient Enrichment Effects on Methylmercury (MeHg)
Nutrient availability—the relative availability of nitrogen (N) and P as well as of S—can
impact MeHg fraction, with low or intermediate levels of nutrients most conducive to
higher MeHg across two very different nutrient gradients. In seven wetlands across a
nutrient gradient in Sweden, total Hg was 113-287 ng/g in soils, MeHg was 3.5-21 ng/g
in soils, and the highest MeHg fraction was 16%. Potential methylation rate constants and
MeHg fractions were highest in wetlands with an intermediate nutrient status (Tjerngren
July 2018
12-80
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
ct al.. 2012). N and P concentrations have also been implicated as contributing to
variation in Hg methylation rate in the Everglades (see Appendix 12.3.4.3).
Sulfur Impacts on Mercury in Wildlife
Mercury has no known beneficial property in living cells. MeHg is bioconcentrated by
living cells, in which it forms strong bonds with thiosulfate and sulfhydryl groups in
organic compounds. It is common for MeHg to be present at progressively higher
concentrations up the trophic level of food chains, as MeHg consumed is remarkably
persistent (Figure 12-13). A recent report by the USGS reviews the known effects of Hg
on wildlife. Fish and birds experience negative physiological and reproductive effects at
Hg levels below the 0.3 ppm tissue level set by the U.S. EPA to protect human health
(Wentz et al.. 2014). There are a number of new studies on the effects of sulfur upon Hg
in wildlife (Figure 12-12).
12-81
DRAFT: Do Not Cite or Quote

-------
r Top
predator fish
rtebrates
Algae and other
inicroorganisnrs
Water
USEPA fish tissue mercury criterion =0,3 parts per million
0.001
Łr 0.0001
E
H 0.00001
0.000001
0.0000001
Note logarithmic scale of y-axis. The top and bottom of each circle represent the range of measured methylmercury concentrations
from streams in Oregon, Wisconsin, and Florida sampled during 2002-2006, and in New York and South Carolina sampled during
2007-2009.
Source: from Wentz et al. (2014).
Figure 12-13 Bioconcentration and biomagnification result in methylmercury
concentrations about 1 million times higher in predator fish than
in stream water.
0.00000001
Increasing trophic level
Forage itsh
My 2018
12-82
DRAFT: Do Not Cite or Quote

-------
Table 12-12 New studies on sulfur (S) impacts upon mercury (Hg) in wildlife.
Type of
Ecosystem
Additions or Load
(kg S/ha/yr)
Biological and
Chemical Effects
Study Site Study Species Reference
River	S addition or
deposition not
reported
[SO42"] in Rio Grande:
pore water: (means)
74-230 mg/L; deep
water: (range)
152-342 mg/L
[SO42"] in Devils
River: pore water:
(means)
74-230 mg/L; deep
water: (range)
15-168 mg/L
M. salmoides muscle Hg
(mg/fish kg) was 38%
higher in DR than RG.
Rio Grande
(RG) and Devils
River (DR)
branches of
Amistad
International
Reservoir, TX
Sediment, pore
water, and
deep water;
Micropterus
salmoides
(largemouth
bass)
Becker et al.
(2011)
Constructed
S load as [SO42 ] in
Fish Hg (Y = log mg
Western Palm
Gambusia
Fena et al.
wetlands
wetland inflow:
Hg/kg G. holbrooki) was
Beach County,
holbrooki
(2014)

39-110 mg/L
negatively correlated with
FL
(eastern



water sulfate (X = log mg

mosquitofish)



SC>42"/L) concentrations

and surface



in three treatment

water



wetlands:





Y = -2.09X + 2.60;





Y = -3.17X + 3.80;





Y = -1.09X- 0.01.



Wetlands
S load as [SO42"] in
Fish Hg (mg Hg/kg or mg
Everglades
Gambusia spp.
Gabriel et al.
and canals
surface water:
Hg/kg/mm for Lepomis)
protection area,
(mosquitofish),
(2014)

0-60 mg/L
increased over 0-1 mg/L
FL
Lepomis spp.



SO42", reached peak

(sunfish), and



levels at 1-12 mg/L

Micropterus



SO42", and decreased

salmoides



over 12-25 mg/L SO42",

(largemouth



to remain flat at SO42"

bass)



concentrations





25-60 mg/L.



Lakes
Not reported
Fish Hg is positively
One
Sander vitreus
Stone et al.


correlated (r2 = 0.928)
impoundment
(walleye)
(2011)


with lake water SO42"
and five natural




(mg/L).
lakes, SD


July 2018
12-83
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Table 12-12 (Continued): New studies on sulfur (S) impacts upon mercury (Hg) in
wildlife.
Type of Additions or Load	Biological and
Ecosystem (kg S/ha/yr)	Chemical Effects	Study Site Study Species Reference
Bog
Ambient S deposition:
5.5 kg/ha/yr (mid
2000s) as quantified
by NADP Station
MN16
Hg levels in Culex spp.
increased 126%.
Culex spp. Hg was 41%
lower in recovery
treatment than under
SO42" treatment, but still
34% higher than control.
S6 peatland,	Pore water, Wasik et al.
bog section,	peat, and	(2012)
Marcell	Culex
Experimental	spp.(mosquito)
Forest, MN	larvae
S addition:
32 kg S/ha/yr
dissolved in pond
water and delivered
by sprinklers (mimics
4x the 1990s
deposition rate)
DR = Devils River; ha = hectare; Hg = mercury; kg = kilogram; L = liter; mg = milligram; NADP = National Atmospheric Deposition
Program; RG = Rio Grande; S = sulfur; S042" = sulfate; yr = year.
S deposition affects the production and net amount of MeHg in water bodies, increasing
accumulation of Hg up the food chain (Table 12-12) but not affecting the rates at which
Hg is transferred between trophic levels. Early work by Bloom et al. (1991) in seepage
lakes of varying pH in the Northern Highland Lake District, WI, considered whether
varying deposition would affect the rates at which MeHg accumulates in biota. The
authors sampled five lakes with common geology but a range of pH (5.1-7.2, including
the experimentally acidified north basin of Little Rock Lake), which the authors ascribed
to varying acidic deposition loads. The fraction of Hg was higher in Perca flavescens,
yellow perch, than in seston, reflecting bioaccumulation of MeHg at higher trophic
levels, but there was no detected effect of pH (and by extension, S deposition load) on
relative bioaccumulation across lakes (Bloom et al.. 1991). In Dickie Lake, Ontario, early
studies documented the biomagnification of Hg in higher trophic levels of the lake.
MeHg concentrations were 0.2 ng/g in sediments; Hg concentrations were 56.5 ng/g in
periphyton, 143 ng/g in P. flavescens, and 703 ng/g inMicropterus spp. [largemouth and
smallmouth bass, (Kerry etal.. 1991)1. Sulfate effects on biomagnification were not
documented in these studies, although sulfate was measured at 6-8 mg/L in lake water,
and microbial sulfate reduction and Hg methylation within the lake sediments were
demonstrated with sediment slurry incubations (Kerry et al.. 1991). There is evidence
from 44 Adirondack lakes sampled in 2003-2004 that the legacy of acidic deposition,
low surface water pH, low ANC, and high aluminum concentrations, all correlated with
increasing Hg concentrations in zooplankton and Hg body burdens in fish and loons (Yu
et al.. 2011). Because these factors did not affect Hg or MeHg concentrations in water,
Yu et al. (2011) concluded that acidic lake conditions enhanced bioaccumulation of
MeHg.
July 2018
12-84
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
Experimental S addition studies demonstrate that S addition increases Hg burdens in
different trophic levels. Little Rock Lake, WI was the site of a multiyear (1985-1990),
progressively acidic S addition study, in which the basins of this lake were separated by a
plastic barrier, allowing the northern half to be experimentally acidified while the
southern half served as a control. Abiotic compartments and biota of the lake were
intensively sampled during the course of the experiment. Aqueous MeHg concentrations
were 2 times higher in the treatment basin, while zooplankton MeHg concentrations
(ng/g) were 1.8 times higher and microseston MeHg concentrations were 2.75 times
higher in the treatment than the control basin (Frost etal.. 1999; Watras and Bloom.
1992). S amendment also affected MeHg in Perca flavescens (yellow perch); MeHg
concentrations were 12-100% higher in 1-year-old yellow perch from the treatment
basin, and the annual mean burden (fig MeHg/fish) was 16-214% higher in the treatment
basin over 5 of the 6 years of the experiment (Frost etal.. 1999; Wiener etal.. 1990).
Decreasing S loads result in decreased Hg burdens in biota. The S addition experiment at
the Marcell peat bog also shows the effects of S addition on Hg burdens in an animal on a
lower trophic level. In the S addition and S recovery experiments at the S6 peat bog,
Marcell Experimental Forest, MN, elevated S increased Hg levels in sampled Culex spp.
(mosquito) larvae. Mosquito larvae are important to consider as a food source because
they are widely consumed by invertebrate predators, fish, amphibians, and birds. In
spring 2009, mosquito larvae Hg was 126% higher in plots that had received
32 kg S/ha/yr addition for 9 years than in mosquito larvae from the control plots (Wasik
ct al.. 2012). The recovery plots received 32 kg S/ha/yr from 2001-2006 and no S
addition after 2006, and when these plots were sampled in 2009, Culex spp. (mosquito)
larvae had Hg levels 34% higher in recovery plots than in control plot mosquitoes,
indicating legacy effects of past S addition. However, recovery plot mosquito Hg
concentrations were 41% lower than in mosquitoes from plots still receiving S additions,
indicating that decreases in S addition will decrease biotic Hg concentrations (Wasik ct
al.. 2012).
In stormwater treatment wetlands draining the EAA in Florida, sulfate concentrations
have been elevated by agricultural runoff, and concentrations ranged from 39-110 mg/L
between 2000 and 2007. Under these very high sulfate surface water concentrations, Hg
levels (mg Hg/kg fish) in bioindicator fish species Gambusia holbrooki (mosquitofish)
were negatively correlated with inflow water sulfate concentrations (Feng et al.. 2014)
because the higher end of the sulfate concentration range reached levels inhibitory to
SRPs. In the Everglades Protection Area, where water sulfate concentrations were lower
and ranged from 0-60 mg/L in surface water, relationships between sulfate and fish Hg
levels were more complicated. Gambusia spp. (mosquitofish), Lepomis spp. (sunfish),
and largemouth bass were sampled annually between 1998 and 2009 (Gabriel et al..
12-85
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
2014).	In all three fishes, Hg tissue concentrations increased between 0-1 mg sulfate/L in
the water, and reached their highest tissue levels between 1 and 12 mg sulfate/L (see
Figure 12-14). Based on these relationships, Gabriel et al. (2014) recommended a sulfate
water standard of 1 mg/L. Two papers challenged this analysis and interpretation of the
sulfate and fish Hg data. In a commentary, Julian et al. (2015) suggested that the
variability in fish Hg levels was too high to correlate with sulfate concentrations, and in
analyses of largemouth bass collected between 2005-2011, Hg concentrations in fish
correlated with alkalinity, pH, and specific conductance as well as sulfate (Julian and Gu.
2015).	Gabriel et al. (2015) responded to Julian's commentary by highlighting the
similarity of fish Hg patterns to documented MeHg water column concentrations in the
Florida Everglades and suggested that sulfate reductions are the most efficient way to
reduce Hg loads in fish because about 60% of the Everglades area has sulfate
concentrations exceeding 1 mg/L (Orem etal.. 2011). Using structural equation
modeling, Pollman (2014) showed that sulfate was second only to periphyton MeHg
concentrations in predicting Hg burdens in mosquitofish in the Everglades, although the
relationships among predictive factors were complex and involved both direct and
indirect effects.
A number of studies document Hg burdens in fish without also quantifying contributions
of S to methylation, but other relationships between S and MeHg documented in the 2008
ISA suggest that the findings of these studies may be relevant (U.S. EPA. 2008a). In the
Alabama River basin around Mobile, AL, there was no relationship across 52 sites in the
watershed between Hg burdens in largemouth bass and MeHg fractions in sediments or
water (Warner et al.. 2005). However, a subset of seven of the sites were downstream of
wetlands, and in these sites the bass Hg burden was positively correlated with watershed
area (r2 = 0.71), indicating that wetlands may contribute to higher Hg burdens in fish as
well as to higher MeHg concentrations in water (Warner et al.. 2005). In a study of Hg
burdens in South Dakota, Stone etal. (2011) analyzed water quality and fish samples
from one impoundment and five natural lakes, and found a positive correlation
(If = 0.928) between water sulfate measurements and Hg in piscivorous Sander vitreus
(walleye). However, sulfate concentrations were not reported, and sulfate measurements
and fish samples were not collected simultaneously, with a gap of up to 3 years between
collection of fish and water samples. In the Amistad International Reservoir, Texas,
elevated MeHg in sediments of one branch of the reservoir correlated with Hg
accumulation in young (age: 0-3 years) largemouth bass. The Devils River area had
elevated sediment MeHg compared to the Rio Grande area of the reservoir (see
Appendix 12.3.5.1). and the length-standardized mean of muscle Hg was 38% higher in
Devils River than in Rio Grande bass (Becker etal.. 2011). Of 138 largemouth bass (age
>3 years) line-caught at the reservoir in April 2007, 77% exceeded the U.S. EPA
recommended human consumption value of 0.3 mg Hg/kg (or 0.3 ppm), and mean fish
12-86
DRAFT: Do Not Cite or Quote

-------
Hg was 0.51 mg/kg fish, although fishing site data for these samples were unavailable,
preventing assessment of connections between sediment MeHg and Hg levels in these
large fish (Becker et al.. 2011).
0.35
0.30 -
0.26 -
¦ running average
D
i J~T~H m
• I	I
i iij i	•
_	I J ~* ~ W\ I
li l» • I ml	rnJ=i
I	®H	• ^ I

-•	hH
rm	1—i—i i i i 11111111 ii 11	1—i—i i i i i 1111111111'	1—i—i i i i 1111111
0.006
e
E
o>
E
0.005
0.004
0.003
0.002
0.001 -
0.000
E
w-
ft.#"
*•%:<
"3i
Ł

o>
<
Surface water sulfate (mg/L)
Surface water sulfate (mg/L)
L = liter; kg = kilogram; mg = milligram; mm = millimeter; THg = total mercury.
Panels A-C show the running average of mercury concentrations in (A) Gambusia spp. (n = 484), (B) Lepomis spp. (n = 2,559), and
(C) Micropterus salmoides (n = 679). Panels D-F show the same data with sulfate transformed to a log scale to better illustrate the
My 2018
12-87
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
peak in fish Hg concentrations that approximately corresponds to sulfate concentrations of 1-10 mg/L. Samples were collected from
12 fish and 12 sulfate stations within the Everglades, between 1998-2009.
Source: Figure 3 in Gabriel et al. (2014).
Figure 12-14 Tissue mercury concentrations as a function of surface water
sulfate concentrations (n = 2,360 surface water samples) in the
Everglades Protection Area.
12.5. Extent and Distribution of Sensitive Ecosystems
Reservoirs in which water levels fluctuate enough to expose sediments may be
particularly susceptible to S deposition effects on MeHg production because mud flats are
directly exposed to the atmosphere and then flooded to create conditions favorable to Hg
methylation by SRPs. Older work in the ELA, Ontario, showed that reservoir creation
leads to a pulse of MeHg production. A small watershed was flooded in 1993, and Hg
and MeHg fluxes were measured over the next decade (St. Louis et al.. 2004). Before
flooding, 90% of the annual total Hg input and 130% of the annual MeHg was retained
by the wetland. In the first 3 years following flooding, 100-170% of the annual total Hg
input was exported annually; and in Years 5-9, 69-79% of the annual total Hg input was
exported annually. MeHg levels were high following flooding, 60-80% in water, and
300-860% of the annual MeHg input was exported annually from the reservoir (St. Louis
et al.. 2004). The reservoir creation also affected Hg in the food web: the level of MeHg
in zooplankton was 7 to 10 times higher in the first 3 years of flooding, and 14 to
30 times higher Years 4-9, than MeHg in preflood zooplankton (St.Louis et al.. 2004).
Researchers stocked finescale dace (current scientific name, Chrosomus neogaeus)
annually and found that the Hg body burden of the fish increased 306% following a year
in the reservoir (St.Louis et al.. 2004). More recently, at the Cottage Grove Reservoir,
OR, elevated MeHg concentrations in the fall outflow (4-13% MeHg fraction compared
to 0.5-7% MeHg in the river flowing into the reservoir) were linked to seasonal water
fluctuations (Ecklev et al.. 2015). The reservoir water level is lowered from September to
February to allow the dam to catch and regulate spring flooding, so 50% of the reservoir
area is exposed mud flats in the winter. Sulfate was below detection limits (0.005 mg
sulfate/g sediment, reported as 5 jxg/g) in permanently inundated reservoir sediments, but
was 0.0011 mg sulfate/g sediment (11 jxg/g) in mudflat sediment. MeHg concentrations
were higher at sediment surfaces than in the sediment profile at seasonally inundated
sites, indicating that MeHg production was occurring there; in contrast, [MeHg] was
constant across sediment depths in the permanently inundated sites, indicating that
July 2018
12-88
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
settling of MeHg into sediments was the only process occurring there (Ecklev et al..
2015).
The National-Scale Assessment of Mercury Risk to Populations with High Consumption
of Self-Caught Freshwater Fish of 2011 (U.S. EPA. 201 lb) compiled fish Hg load data
on a national scale, which can be considered a county-wide sampling of ecosystems
where Hg loads in fish are high. The assessment did not directly quantify or analyze
sulfate effects, but it considered patterns of Hg deposition and Hg concentrations in
freshwater fish caught between 2005-2009 by state regulatory agencies, and modeled the
extent to which reductions in Hg deposition would affect subsidence-level human
consumers of fish from U.S. water bodies. Fish collections showed that there were
elevated fish Hg levels in water bodies across the country, although high fish Hg
concentrations were more common in the eastern U.S. [see Figure 12-15 (U.S. EPA.
2011b)l. In this study, 5% of total Hg deposition was attributed to coal and oil-fired
electric utility steam generating units (EGUs), but EGUs contributed 9% to fish tissue
levels of Hg.
The 2014 USGS's report on mercury in streams (Wentz et al.. 2014) suggests that eastern
water bodies, watersheds with high wetland cover, and watersheds with low population
densities are linked to higher MeHg concentrations in streams and in stream biota.
Fish Tissue Mercury Data
(HUC12-level 75th percentile values, ppm)
~	0.000000 - 0.109143
o	0.109144 - 0.195000
o	0.195001 - 0.300000
o	0.300001 - 0.479000
*	0.479001 - 6.605000
My 2018
12-89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
HUC12 = 12-digit hydrologic unit code.
Source: U.S. EPA (2011b).
Figure 12-15 Fish mercury concentrations across the U.S.
12.6. Critical Loads
There are currently no critical loads for SOx deposition related to the nonacidifying
effects described in this appendix. There are European critical loads for Hg deposition
and Hg effects upon human health and ecosystems. A critical load of 1 |_ig Hg/L in soil
drainage water was set to protect drinking water (Hcttclingh et al.. 2015). A critical load
of 0.3 mg Hg/kg fresh weight fish tissue was set to protect human health. A critical load
to protect soil invertebtates and soil microbial process of 0.5 mg Hg/kg organic material
was set for the humus layer of forest soils (Hcttclingh et al.. 2015). This source did not
report critical loads in terms of areal deposition rates, but did map exceedances based on
soil, water, and fish tissue samples, and found that about 59% of European natural areas
received Hg deposition in exceedance of these critical loads (Hcttclingh et al.. 2015).
12.7. Summary of Nonacidifying Sulfur Effects
Sulfate deposition has a number of effects beyond acidification in ecosystems. In the
2008 ISA, the evidence was sufficient to infer a causal relationship between S deposition
and increased methylation of Hg in aquatic environments where the value of other factors
was within adequate range for methylation. The 2008 ISA surveyed literature from
approximately 1980 to 2008 and described the qualitative relationships between sulfate
deposition and a number of ecological endpoints, including altered S cycling, sulfide
phytotoxicity, internal eutrophication of aquatic systems, altered methane emissions,
increased Hg methylation, and increased Hg loading in animals, particularly fish. New
evidence from wetland and freshwater aquatic ecosystems strengthens and extends the
causal findings of the 2008 ISA regarding nonacidifying sulfur effects, and provides the
basis for a new causal determination. Table 12-13 summarizes recent thresholds and
quantitative relationships describing effects of sulfate deposition. This new information
strengthens the weight of evidence presented in the 2008 ISA, and consistent with the
2008 causal statement, the body of evidence is sufficient to infer a causal relationship
between S deposition and the alteration of Hg methylation in surface water,
sediment, and soils in wetland and freshwater ecosystems. In addition, recent research
July 2018
12-90
DRAFT: Do Not Cite or Quote

-------
1	demonstrates sulfide phytotoxicity under current conditions in North American wetlands,
2	consistent with the 2008 ISA's summary of sulfide phytotoxicity in European wetland
3	and freshwater ecosystems. The body of evidence is sufficient to infer a causal
4	relationship between S deposition and changes in biota due to sulfide phytotoxicity
5	including alteration of growth and productivity, species physiology, species richness,
6	community composition, and biodiversity in wetland and freshwater ecosystems.
7
Table 12-13 Summary of quantitative effects of nonacidifying sulfur enrichment.
Section of
Nonacidifying Sulfur
Effects Appendix
Threshold, Critical Level, or Quantitative Relationship
Reference
12.2.1
2.5-7.9 times higher sulfate reduction in lake with 22.7 mg
S042"/L than in lake with 9.0 mg S042"/L
Kleeberq etal. (2016)
12.2.3
34.1 mg S2"/L for plant root nutrient uptake
Koch etal. (1990)
12.2.3
48-50 mg S042"/L as a threshold to protect Potamogeton spp.
and Utricularia vulgaris
Vermaat et al. (2016):
Smolders et al. (2003)
12.2.3
0.3-29.5 mg S2"/L for altered growth, productivity, physiology,
or mortality of 16 freshwater wetland emergent plant and
aquatic submerged macrophyte species native to North
America
Lamers et al. (2013)
12.2.3
250 mg S042"/L U.S. EPA aesthetic secondary water quality
standard for human consumption
Pastor et al. (2017)
12.2.3
<7.5 mg/L sulfide to preserve growth rate of Cladium
jamaicense, sawgrass, in Everglades
Li et al. (2009)
12.2.3
0.165 mg sulfide/L to protect Zizania palustris, wild rice
MPCA (2015a. 2015b)
12.3.2
Sulfate reduction rates increase 0.35 mg (sulfate
reduced)/L/day with a 1 mg/L increase in sulfate addition
Kerrvetal. (1991)
12.3.3.1
11.1 mg SC>42"/L added for peak MeHg production in incubation
of sediments from Quabbin Reservoir, MA
Gilmour et al. (1992)
12.3.3.1
5.8-23 mg SC>42"/L added to change incubation of sediments
from net demethylating to net demethylating, from Little Rock
Lake, Wl
Gilmour and Riedel
(1995)
12.3.3.1
190 mg SC>42"/L added to increase MeHg production in
incubation of peat from Sunday Lake, NY
Yu etal. (2010)
12.3.3.1
96.1	mg SC>42"/L increases methylation rates in incubations of
sediment from South River, VA (ambient concentrations of
19.2	mg S042"/L)
Yu etal. (2012)
July 2018
12-91
DRAFT: Do Not Cite or Quote

-------
Table 12-13 (Continued): Summary of quantitative effects of nonacidifying sulfur
enrichment.
Section of
Nonacidifying Sulfur
Effects Appendix
Threshold, Critical Level, or Quantitative Relationship
Reference
12.3.3.5
<3.6 mg C/L as dissolved organic carbon, microbial sulfate
reduction did not occur
3.6-7.6 mg C/L of DOC, microbial sulfate reduction increases
linearly with C increase
>7.6 mg C/L, microbial sulfate reduction rate does not change
Watras et al. (2006)
12.3.3.5
20% organic material in sediments for peak MeHg
concentrations in Adirondacks, NY watersheds
Yu et al. (2010)
12.3.3.5
2.65 ng/L increase in MeHg for every 10% increase in
percentage organic matter
Hoaaarth et al. (2015)
12.3.3.5
1 ng MeHg/L increase for each increase of 0.048 mg
hydrophobic organic acid fraction of DOC/L
Hall et al. (2008)
12.3.3.5
0.34 ng Hg/L increase in Hg exported in streams for each 1 mg
C/L increase in DOC from Northeast watersheds
Dittman etal. (2010)
12.3.3.7
1.0 mg N-N03"/L in hypolimnion to prevent mercury methylation
in Onondaga Lake, NY
Matthews etal. (2013)
12.3.4.1
Addition of 4.8 mg S042"/L increases methylation rate in Little
Rock Lake, Wl (ambient water concentration, 2.4 mg S042"/L)
Gilmour and Riedel
(1995)
12.3.4.1
10.8 ng MeHg/L increase for each 1 mg/L increase in H2S in
lake water
Watras et al. (2006)
12.3.4.1
14.4 mg total lake MeHg increase for every 1 kg/ha increase in
SOx deposition
Watras and Morrison
(2008)
12.3.4.1
14 kg S/ha increased MeHg concentrations in peat mesocosm
Branfireun etal. (1999)
12.3.4.1
<8.3 kg S/ha addition for pore water MeHg in peatlands
Mitchell etal. (2008a)
12.3.4.1
<32 kg S/ha/yrto control pore water MeHg concentrations and
fraction of total Hg in peatlands
Jeremiason et al. (2006)
12.3.4.1
<20 kg S/ha/yr to control pore water total Hg concentrations and
MeHg concentrations in peatlands
Berqman et al. (2012)
12.3.4.3
1 mg/L sulfate in surface water to keep MeHg concentrations
low in Everglades surface water
Corrales et al. (2011)
12.3.4.3
>20 mg S042"/L inhibits Hg methylation in Everglades due to
sulfide accumulation
Orem et al. (2011)
12.3.4.3
>39 mg S042"/L inhibits Hg methylation in Everglades due to
sulfide accumulation
Zhenq etal. (2013)
12.3.4.3
50 ng total Hg/L, the U.S. EPA water quality criterion
Atoers et al. (2014)
July 2018
12-92
DRAFT: Do Not Cite or Quote

-------
Table 12-13 (Continued): Summary of quantitative effects of nonacidifying sulfur
enrichment.
Section of
Nonacidifying Sulfur
Effects Appendix
Threshold, Critical Level, or Quantitative Relationship
Reference
12.3.5.1
<10 kg S/ha/yr, decreases in S deposition result in declining fish
Hg concentrations on a decadal time scale in Isle Royale lakes
Drevnick et al. (2007)
12.3.5.1
MeHg concentrations increased 0.037 ng/L for each 1 mg/L
increase in sulfate in pore water in peatlands
Mitchell et al. (2009)
12.3.5.1
0.06 ng increase in MeHg per gram of sediment for every
1 mg/L increase in sulfate pore water concentration in South
River, VA
Yu et al. (2012)
12.3.5.1
0.11 ng/L increase in stream MeHg concentration for each mg/L
sulfate decrease in Archer Creek, NY
Selvendiran et al.
(2008a)
12.3.5.1
65% higher MeHg in Devils River branch (SO42" 223 mg/L) than
in Rio Grande branch (SC>42~2230 mg/L) of Amistad Reservoir,
TX
Becker et al. (2011)
12.3.5.1
MeHg increases with S reduction in wetland where average
SO42" 25 mg/L, but not in wetland where SO42" 2500 mg/L
Johnson et al. (2016b)
12.3.5.2
11.9 kg S deposition/ha/yr increases Hg levels in largemouth
bass compared to ecoregions where deposition was 7.7-8.1 kg
S/ha/yr in Texas
Drenneretal. (2011)
12.3.5.2
MeHg decreased in surface water and in Perca flavescens
when S deposition decreased from 6.44 kg S/ha/yr (1998) to
3.35 kg S/ha/yr (2012) in two lakes in Voyageurs NP
Briqham et al. (2014)
12.3.5.3
When conductivity and aromatic DOC remain constant, surface
water MeHg increases 0.3 ng/L for each mg/L increase in
sulfate concentrations
Hall et al. (2009a)
12.4
<32 kg S/ha/yr, for Hg load in larval Culex spp. (mosquitoes) in
peatlands
Wasiketal. (2012)
12.4
1 mg/L sulfate in surface water to keep Hg concentrations low in
fish: Gambusia spp., Lepomis spp., and Micropterus salmoides,
in the Everglades
Gabriel et al. (2014)
C = carbon; DOC = dissolved organic carbon; H2S = hydrogen sulfide; ha = hectare; Hg = mercury; kg = kilogram; L = liter;
MeHg = methylmercury; mg = milligram; ng = nanogram; S = sulfur; yr = year.
1
2	12.7.1. Terrestrial Sulfur Cycling
3	The 2008 ISA noted that S is an essential plant nutrient and that S deposition can affect
4	plant protein synthesis by affecting S availability for S-containing amino acids as well by
5	affecting N uptake. Sulfate can be taken up directly by plant roots as well as by soil
July 2018
12-93
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
microbes, or it can adsorb to soil particles. The 2008 ISA reported that watersheds in the
Southeast, which retained more S than was deposited during high S loading in the 1980s,
received decreasing S deposition over the 1990s but continued to export historically
deposited sulfate in streams. In the Northeast, budgets of sites from the 1980s indicated
that high S deposition caused high sulfate export in streams. On a national level, S
deposition resulted in increased S content in organic matter in soils, and studies suggested
that mineralization of this stored S may contribute to elevated sulfate leaching for
decades after reductions in S deposition. More recent research confirms that watersheds
in the Northeast continue to leach more S than they currently receive from S deposition,
as much as 24-45% more S than the S deposition load (Mitchell et al.. 2011).
Aquatic Sulfur Cycling
The 2008 ISA reported that nonacidifying effects of S in freshwater systems will be
affected by water residence times, with larger lakes and beaver ponds providing ample
opportunity for SRPs to reduce sulfate and generate acid-neutralizing capacity (ANC).
Early work showed that a 1 mg/L increase in sulfate concentration in Dickie Lake,
Ontario, increased S reduction rates 0.35 mg S/L/day (Kerry et al.. 1991).
The 2008 ISA reported the role of wetlands in removing sulfate from surface water and
storing it in sediments or biomass, but it also highlighted the way drought reverses this
capacity to make wetlands sources of S to downstream waters. Specifically, studies
reviewed in the 2008 ISA demonstrated that under drought conditions, sulfide in
previously saturated moss or sediments is exposed to atmospheric oxygen and reoxidized
to sulfate. This resulted in high sulfate concentrations in surface water when the water
table rose to its former level, and MAGIC modeling showed that variable water levels
due to climate change-induced droughts would delay ANC recovery in the Plastic Lake
Watershed, NY (Aherne et al.. 2004). In Little Rock Lake, WI, water level fluctuation
due to drought added an additional 5 kg S/ha/yr from internal ecosystem S stores to the
lake (Watras and Morrison. 2008). Recent work confirmed that periods of drought result
in elevated water sulfate concentrations during the recovery period following the drought
(Wasiketal.. 2012).
Sulfide Toxicity
In aquatic systems or in saturated soils, SRPs convert sulfate into sulfide, which inhibits
nutrient uptake in freshwater aquatic and wetland plant species. The 2008 ISA showed
that sulfide toxicity reduced biomass of wetland plants and aquatic macrophytes in
12-94
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
mesocosms under aquatic S concentrations higher than those that occur in U.S. regions
with high S deposition. Recent research has shown sulfide phytotoxicity occurs at
ambient aquatic S concentrations within multiple ecosystems in the U.S. (Figure 12-16).
Sulfide decreases total plant cover and cover of dominant species in a New York fen
(Simkin et al.. 2013) and decreases the growth rate of Everglades, FL, keystone species
Cladium jamcticense (sawgrass) at surface water concentrations of 7.5 mg sulfide/L (Li et
al.. 2009). The state of Minnesota is also working on a sulfide standard to protect the
economically and culturally important Zizania palustris (wild rice) and has proposed a
water standard of 0.165 mg sulfide/L to protect the species (MPCA. 2015a). Additionally,
a review by Lamers et al. (2013) reported experimentally-determined sulfide toxicity
values between 0.3-29.5 mg S27L for 16 North American species (see Table 12-2). and
there are sulfate tolerance values for North American fish and macroinvertebrates based
on USGS's NAWQA. This new information shows that sulfide toxicity occurs in North
American wetlands under current deposition conditions, and the body of evidence is
sufficient to infer a causal relationship between S deposition and changes in biota
due to sulfide phytotoxicity including alteration of growth and productivity, species
physiology, species richness, community composition, and biodiversity in wetland
and freshwater ecosystems.
Water quality thresholds for non-acidifying S effects
Increase in MeHg concentrations in surface water of Everglades
1 mg sulfate/L |ncrease in mosquitofish, sunfish, and largemouth bass Hg load in Everglades
7.5 mg sulfide/L Decrease in sawgrass
growth rate in Everglades
| 0.165 mg sulfide/L Decline in populations of wild rice in MN
0	2	4	6	8
mg/L
L = liter; Hg = mercury; MeHg = methylmercury; mg = milligram; S = sulfide.
July 2018
12-95
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
Figure 12-16 Thresholds of sulfate or sulfide concentrations in water, which
cause biological and chemical effects in ecosystems.
12.7.4. Internal Eutrophication
The 2008 ISA described the contribution of S deposition to internal eutrophication in
aquatic systems. In wetland and lake waters, sulfate is reduced to sulfide, which reacts
with Fe to form insoluble iron sulfide complexes. In many ecosystems, the iron in this
reaction is provided by FePC>4, which releases phosphorus as it forms FeSxthus
contributing to downstream eutrophication. More recently, internal eutrophication caused
by sulfate addition was observed in mesocosms of samples collected from Lake Moshui,
China (Yu et al.. 2015).
12.7.5. Effects on Methane Production
Sulfate deposition can shift microbial community interactions, resulting in lower methane
emissions. The 2008 ISA documented the suppression of methane emissions in wetland
soils by sulfate addition in several studies and noted that 15 kg S/ha/yr (the lowest S load
in experimental treatments) suppressed methane emissions to the same extent as higher S
loads (Gauci et al.. 2004). There is no recent research that relates SOx deposition to
wetland methane emissions, but recent research has confirmed the underlying microbial
mechanisms of this process. Sulfate deposition increases the abundance or metabolic
activity of SRPs. In many freshwater water bodies, including wetlands, low primary
productivity limits microbially available, labile C during the growing season when higher
temperatures favor microbial activity. Recent research documents competition among
microbial guilds for labile C, and when SRPs are favored, competing methanogens
decline, resulting in suppressed methane emissions (Bae et al.. 2015; He et al.. 2015;
Paulo et al.. 2015). However, there is also evidence of methanogens forming syntrophic
associations with SRPs (Paraniapc et al.. 2017). suggesting that these groups may not
always directly compete in all ecosystems.
12.7.6. The Role of Microbes in Mercury Methylation
The 2008 ISA identified sulfur-reducing bacteria as responsible for Hg methylation, and
identified anoxic wetland and lake bottom sediments as their location within watersheds.
July 2018	12-96	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
More recent research confirms that SRB in these locations methylate Hg, but also shows
that the ability to methylate Hg is more broadly distributed in the prokaryotic taxa and
across the environment.
Recent research suggests that the hgcAB gene pair confers the ability to methylate
inorganic Hg. HgcAB has been sequenced in lineages within bacteria and archaea, which
is why this document refers to mercury methylators as SRPs rather than the SRB
described in the previous ISA (Gilmour et al.. 2013). Experimental work has shown that
MeHg production rates vary among prokaryotic strains (Shao et al.. 2012). but also has
confirmed that sulfate addition causes Hg methylation in the Everglades (Goni-Urriza et
al.. 2015). as abundant syntrophs (heterotrophs dependent on the metabolic byproducts of
other microbes) were responsible for both Hg methylation and sulfate reduction (Bac et
al.. 2014).
Recent research shows that the microbial communities responsible for Hg methylation
are more widely distributed than in the freshwater lake or wetland bottom sediments
described in the 2008 ISA (Figure 12-17). There is evidence from Sweden, Finland, and
the U.S. that disturbances in terrestrial forests stimulate mercury methylation in soils and
MeHg export to surfacewaters (Kronbcrg et al.. 2016; Poulin et al.. 2016; Ukonmaanaho
et al.. 2016). The Sphagnum (moss) mat of wetlands was the most efficient retention site
of deposited Hg in a forested watershed in the Adirondacks (Selvendiran et al.. 2008b). as
well as an important location of mercury methylation within peat wetlands in New York
and Wisconsin (Yu et al.. 2010; Creswell et al.. 2008; Selvendiran et al.. 2008b).
Experiments using samples from rivers or streams in Virginia, Minnesota, or Germany
showed that higher sulfate concentrations increased methylation rates or [MeHg] (Frohnc
et al.. 2012; Yu et al.. 2012; Tsui et al.. 2008). with a 0.06 ng/L increase in MeHg per
gram of sediment for every 1 mg/L increase in sulfate pore water concentration in South
River, VA (Yu et al.. 2012). New research shows that SRPs actively methylate Hg within
periphyton, the aquatic biofilms attached to substrate or macrophytes in oxic freshwater
environments (Correia etal.. 2012; Achaetal.. 2011). MeHg production has also been
documented in estuarine and marine sediments of the Chesapeake Bay (Hollweg et al..
2009) and in the marine water columns sampled at 20 to 327 m depths in the Canadian
Arctic Archipelago (Lehnherr et al.. 2011).
12-97
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
10
8
6
SK lakes (ng/L)*
5 -2
o
"SB 2
c
4
MN peatlands
~ (ng/L)
-4
-6
-8
-10
0
5
10	15
Surface water sulfate (mg/L)
20
25
C = carbon; cm = centimeter; DOM = dissolved organic matter; g = gram; L = liter; MeHg = methylmercury; mg = milligram;
mS = millisiemens; ng = nanogram; SK = Saskatchewan; S042- = sulfate.
Note: A study of prairie lakes and wetlands in Saskatchewan found that conductivity, aromatic DOM, and sulfate concentrations
correlate with water MeHg (ng MeHg/L) = -2.94 - 0.44 (conductivity, mS/cm) + 0.303 (mg S0427L) - 0.268 (aromatic fraction of
DOM, L/mg C/cm). Simple linear regressions relate MeHg to water sulfate concentrations in Minnesota peatlands (water
ng MeHg/L = 0.58 + 0.037 mg S042"/L) and in South River, VA (riverbed ng MeHg/g sediment = -0.08 + 0.059 mg S0427L).
Figure 12-17 Linear relationships between sulfate and methylmercury
concentrations in published studies.
As a microbial process, Hg methylation is determined not just by sulfate and Hg
concentrations, but by other environmental and nutritional requirements of SRPs: pH,
temperature, and water quality parameters, and carbon supply. The 2008 ISA identified
pH and dissolved organic carbon (DOC) to be major controls on Hg production, with low
pH and moderately high DOC correlating with high fish MeHg levels in lakes. The 2008
ISA also identified wetland area or density as an important determinant of MeHg
bioaccumulation and export in watersheds. Recent research presented in Appendix 12.3.3
evaluates temperature, total Hg concentration, pH, organic matter in water and sediments,
iron, and nitrate for their influence on Hg methylation rates.
The 2008 ISA identified organic C as an important control on microbial sulfate reduction
and Hg methylation, and recent research has quantified this relationship, hi Little Rock
Lake, WI, DOC concentrations had to be at least 3.6 mg C/L to allow microbial sulfate
reduction, and microbial sulfate reduction increased linearly with an increase of C only
when DOC was in the range of 3.6-7.6 mg C/L (Watras et al.. 2006). Across prairie
My 2018
12-98
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
potholes in Saskatchewan, surface water [MeHg] increased 2.65 ng/L for every 10%
increase in percentage organic matter in underlying sediment (Hoggarth et al.. 2015). In
wetlands, rivers, and lakes of the Mississippi River delta, LA, there was a 1 ng MeHg/L
increase in surface water for each additional 0.048 mg hydrophobic organic acid fraction
(aromatic carbon typical of peat) of DOC/L (Hall et al.. 2008).
Impacts of Sulfur upon Mercury Cycling
Based on experimental evidence, the 2008 ISA identified maximum MeHg production
occurring at sulfate concentrations of 200-400 (j,eq sulfate/L (9.6-19.2 mg SO42 /L) and
noted that waters under S deposition in the U.S. have sulfate concentrations in the range
of 60-200 (j,eq sulfate/L (2.9-9.6 mg SC>427L). More recent research shows that
detectable mercury methylation and bioaccumulation occurs at water sulfate
concentrations >1 mg/L in the Everglades (Gabriel et al.. 2014; Corrales et al.. 2011) and
at ambient S deposition and water sulfate concentrations within the U.S. There was a
positive relationship between sulfate reduction and concentrations of MeHg in the
hypolimnion of Little Rock Lake, WI, with a 10.8 ng MeHg/L increase for each mg S/L
reduced ("Watras et al.. 2006). There was also a positive relationship at this site between
total lake MeHg and annual SOx deposition, with a 14.4 mg MeHg/lake increase for
every 1 kg S/ha/yr increase in SOx deposition (Watras and Morrison. 2008).
There are multiple lines of evidence of a positive relationship between sulfate surface
water concentrations and MeHg concentration or production in multiple freshwater
systems. Both experimental manipulations and observational studies show that higher
sulfate concentrations in peat wetlands increase MeHg concentrations (Figure 12-18).
Experimental addition of sulfate to Bog Lake Fen in Minnesota showed that 8.3 kg S/ha
increased pore water [MeHg] (Mitchell et al.. 2008a). and 32 kg S/ha/yr increased pore
water [MeHg] as well as MeHg fraction of total Hg, or %MeHg (Jeremiason et al.. 2006).
In other bogs sampled in the same region of Minnesota and Ontario, [MeHg] in water
increased 0.037 ng/L for each 1 mg sulfate/L increase (Mitchell et al.. 2009). These
results were confirmed in a peatland S addition experiment in Sweden, which found that
20 kg S/ha/yr addition to the peat increased [MeHg] (Akerblom et al.. 2013; Bergman et
al.. 2012).
Recent work also established positive relationships between sulfate and MeHg in
ecosystems other than peat wetlands. In prairie potholes in the West (Hoggarth et al..
2015). with surface water MeHg increasing 0.3 ng/L for each additional mg sulfate/L
where conductivity and aromatic DOC remain constant (Hall et al.. 2009a). In a
freshwater marsh in the Everglades, water [MeHg] concentrations increased when sulfate
12-99
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
concentrations exceeded 1 mg sulfate/L (Corrales et al.. 2011). which is also an important
threshold value for fish in the Everglades (see below). There are also recent
ecosystem-scale studies that show that sulfate reduction and MeHg can be correlated at
the landscape level. At the watershed level in the Adirondacks, [MeHg] in stream
outflows was negatively correlated with sulfate concentration (Selvendiran et al.. 2009).
with a 0.11 ng/L increase in stream [MeHg] for each mg sulfate/L removed from surface
water (Selvendiran et al.. 2008a).
Deposition thresholds for non-acidifying S effects
I 32 Increase in porewater %MeHg in MN bog
Increase in mosquito Hg load in MN bog
20 Increase in porewater [Hg] in MN bog
11.7 Increase in largemouth bass Hg load in TX reservoirs
5.3 Increase in porewater [MeHg] in MN bog
10	15	20	25	30	35
kg S/ha/yr
ha = hectare; Hg = mercury; kg = kilogram; MeHg = methylmercury; S = sulfur; yr = year.
Figure 12-18 Thresholds of sulfate addition or deposition from published
studies which affect chemical or biological changes in
ecosystems.
12.7.8. Sensitive Ecosystems
The 2008 ISA identified ecosystems in the Northeast as particularly sensitive to Hg
methylation in response to S deposition, because many watersheds there have abundant
wetlands and freshwater water bodies with high DOC and low pH. Recent analyses
July 2018
12-100
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
confirm that ecosystems east of the Mississippi tend to have higher Hg concentrations in
fish ("Wentz et al.. 2014; U.S. EPA. 2011b'). Recent papers suggest that reservoirs with
high water-level fluctuation or high algal productivity also have high methylmercury
production or bioaccumulation within fishes (Ecklev et al.. 2017; Noh et al.. 2016;
Ecklev et al.. 2015; Becker etal.. 2011). and that at the ecoregion level, reservoir fish Hg
concentrations correlate with both Hg and S deposition (Drenner et al.. 201IV
Recent research in the Yolo Bypass Wildlife Area, CA, indicates that Hg methylation
occurs in both permanent and agricultural wetlands and was higher in the agricultural
wetlands where rice is grown (Alpers et al.. 2014; Marvin-Dipasquale et al.. 2014). Hg
methylation in this system was associated with Mn reduction, and then to a lesser extent
with Fe reduction, and was only marginally associated with S reduction; sulfate addition
did not affect Hg methylation (Alpers et al.. 2014; Marvin-Dipasquale et al.. 2014).
However, MeHg produced in the sediments where rice grew was highly correlated with
MeHg concentrations in rice seeds ("Windham-Mvers et al.. 2014a). suggesting a route to
human exposure, and there were higher MeHg burdens in several animal species
collected from agricultural wetlands than from permanent marshes (see below).
Mercury Effects on Animal Species
Mercury is a developmental, neurological, endocrine, and reproductive toxin across
animal species. It is transformed by SRPs from inorganic Hg into methylmercury
(MeHg), which forms strong complexes with thiosulfate and sulfhydryl groups of organic
molecules. This bound MeHg accumulates in biota in successively higher concentrations
at ascending trophic levels. The 2008 ISA documented Hg accumulation in four turtle
species, songbirds, insectivorous passerines, and the common loon (Gavia immer). A
recent report by the USGS notes that Hg accumulation has also been documented in
insectivorous songbirds and bats (Wentz et al.. 2014). In recent research conducted in the
Yolo Bypass, CA, higher water [MeHg] in agricultural rice-producing wetlands than in
permanent marshes led to higher MeHg concentrations in mosquitofish (Gambusia
affinis) and Mississippi silversides (Menidia audens) from the agricultural wetland than
from the permanent marshes (Ackerman and Eagles-Smith. 2010).
The 2008 ISA reported that 23 states had issued fish advisories by 2007 in response to the
U.S. EPA's fish tissue criterion set to protect human health of 0.3 |ig MeHg/g fish, or
0.3 ppm. The 2008 ISA reported on the negative impacts of Hg on development,
morphology, survival, or reproduction in the following fish species: walleye (Stizostedion
vitreum), grayling (Thymallus thymallus), mummichog (Fundulus heteroclitus), rainbow
trout (Oncorhynchus mykiss), fathead minnows (Pimephales promelas), and zebrafish
12-101
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
(Danio rerio). However, a recent report on Hg in streams of the U.S. by the USGS
summarizes current research that birds, fish, and fish-eating wildlife experience negative
effects of Hg at lower concentrations than the 0.3 ppm criteria ("Wentz et al.. 2014).
The 2008 ISA reviewed two studies that considered the link between S deposition and Hg
levels in fish (Drevnick et al.. 2007; Hrabik and Watras. 2002); both studies showed that
decreases in S deposition resulted in decreases in fish MeHg levels. Recent research
supports this finding in Voyageurs National Park (a Class I area) in Ryan Lake between
1998-2012 (Brigham et al.. 2014). when decreasing S deposition correlated with
decreasing fish Hg. In addition, a survey of fish caught in Texas reservoirs found that fish
in the highest deposition region (11.7 kg S/ha/yr) had significantly higher mean Hg levels
than fish from other regions (Drenner et al.. 2011). Experimental S addition to the
Marcell peat bog in Minnesota demonstrated that 32 kg S/ha/yr increased the Hg
concentrations in larval Culex spp. (mosquitoes), which are an important food source for
both aquatic and terrestrial species (Wasik et al.. 2012).
In addition to the studies that consider S deposition, there are recent studies that consider
sulfate concentrations in water in relation to fish Hg concentrations. A study of fish from
six lakes in South Dakota found a positive correlation between sulfate water
concentrations and walleye Hg concentrations (Stone etal.. 2011). The marshes of the
Everglades receive high S loading as agricultural runoff, and recent analysis of Hg loads
in mosquitofish, sunfish (Lepomis spp.), and largemouth bass (Micropterus salmoides)
collected from 1998-2009 showed that Hg levels in fish were highest when sulfate
concentrations were between 1 and 12 mg sulfate/L; the researchers proposed 1 mg
sulfate/L as a water standard (Gabriel et al.. 2014).
12-102
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
SUPPLEMENTAL MATERIAL
MERCURY CYCLING
This supplemental material describes the deposition and cycling of mercury within North
American ecosystems. Connections to sulfur cycling are emphasized where such
information is available. This supplemental material provides background information
and context for the discussion of nonacidifying sulfur stimulation of mercury methylation
in wetland and aquatic ecosystems (Appendix 12).
Transfer of Mercury from the Atmosphere to Terrestrial Ecosystems
Mercury enters most natural terrestrial and aquatic ecosystems in the U.S. via
atmospheric deposition, although both active and historic mining and industrial sites have
contaminated soil and water in many ecosystems. Sources and atmospheric processes
governing Hg deposition are covered more exhaustively elsewhere ("Wentz et al.. 2014;
Pirrone et al.. 2010). For the purposes of this section, we will focus on the deposition of
reactive Hg, which is deposited into ecosystems in the form of free Hg(II),
particulate-bound Hg, or as MeHg. In a jack pine/birch forest in the ELA sampled in
1998-1999, wet deposition was 71,000,000 ng/ha/yr (reported as 71 mg total Hg/ha/yr),
with dry deposition estimated at 9,000,000 ng/ha/yr [reported as 9 mg/ha/yr in (St Louis
etal.. 2001)1.
Foliage is an important Hg pool in both forested and wetland ecosystems because it can
directly absorb Hg deposition as well as store Hg taken up by plant roots. In the Arbutus
Lake watershed in the Adirondack Mountains, Hg in the AI mis and Betula overstory was
3,080 ng Hg/m2 (3.08 |ig Hg/m2), and was 9,480 ng/m2 (9.48 (ig/m2) in the Sphagnum
mat of the wetlands (Selvendiran et al.. 2008b). In a 3-year-long Hg tracer isotope annual
addition experiment at ELA, upland canopy and ground vegetation retained 40% of
added Hg. In a wetland ecosystem, vegetation retained 80% of added Hg (Harris et al..
2007). Although the wetland ecosystems exported Hg and MeHg to an adjacent lake in
the watershed during the course of the study, the Hg was from sources existing before the
start of the experiment, indicating that Hg storage in wetland vegetation may have a
residence time greater than 1 year (Harris et al.. 2007).
Foliage also represents an important flux of Hg; following senescence, leaf litter becomes
the substrate for microbial decomposition, and Hg moves out of plant biomass and into
microbial biomass, soil, or water. Flux of total Hg from litter in the ELA, Ontario, was
twice that of wet deposition in litter from upland trees and was of a similar magnitude to
July 2018
12-103
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
deposition in litter from wetland shrubs and trees [72,000,000 ng/ha/yr or 72 mg
Hg/ha/yr, and 88,000,000 ng/ha/yr or 88 mg/ha/yr; (St Louis et al.. 2001)1. The organic
horizon of the soil, where litter and microbial communities are concentrated, was a major
sink of Hg within the forested and wetland ecosystems for the 20 years preceding this
study, with annual Hg accumulation rates of 130,000,000 ng Hg/ha/yr (reported as
130 mg Hg/ha/yr) for ridgetop soils (160% of total annual deposition),
200,000,000 ng/ha/yr (200 mg/ha/yr) for upland forest (250% of deposition), and
590,000,000 ng/ha/yr (590 mg/ha/yr) in wet soils supporting Sphagnum mosses
[7.4 times annual deposition; (St Louis et al.. 2001)1. In forest catchments, much of the
Hg deposition is retained in soil (St Louis et al.. 2001; St. Louis et al.. 1996). as
confirmed by the Hg tracer isotope addition experiment in ELA, which found that after
3 years of Hg addition, 58% of the added Hg was retained in soil in the upland forest
catchment (Harris et al.. 2007). The length of residence within the soil for Hg likely
depends on the same factors that control the storage of organic matter, such as soil
structure, temperature, precipitation, and timing of disturbances, particularly of wildfire.
Transfer of Mercury from Terrestrial to Aquatic Ecosystems
When Hg is released from plant tissue, detritus, or soil particles into surface water or the
soil solution, it becomes available for transport into aquatic systems. In five boreal forest
catchments in the ELA, Ontario, measured during the period 1990-1993, the flux of total
Hg from the forest into water bodies was estimated to be
14,000,000-21,000,000 ng/ha/yr (reported as 14-21 mg/ha/yr 28-42% of annual Hg
input), in response to approximately 47,000,000 ng/ha/yr (47 mg/ha/yr) wet + dry
deposition of Hg, and estimated contributions from weathering of granite of as much as
3,000,000 ng/ha/yr [reported as 3 mg/ha/yr in St. Louis et al. (1996)1. On Isle Royale, a
large island in Lake Superior and a Class I Area, researchers estimated that lakes received
one-third of their annual Hg load directly from atmospheric deposition, and the remaining
two-thirds of annual Hg load was mobilized from soils in the watersheds (Drevnick et al..
2007). In aquatic environments, specialized microbes (SRPs; as well as iron-reducing
bacteria) in anoxic zones can methylate inorganic Hg, transforming it into the toxic,
persistent MeHg.
July 2018
12-104
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Transfer of Mercury from Atmosphere to Aquatic Ecosystems
Total Hg deposited directly to lakes is rapidly incorporated into Hg cycling in abiotic and
biotic pools. In a Hg isotope addition experiment at the ELA, Ontario, a spike of isotope
202Hg was added directly to the lake annually over the course of 3 years. The 202Hg was
detected in the anoxic hypolimnion within days of addition, and in bottom sediments of
the lake within 4 weeks (Harris et al.. 2007). and methylated 202Hg was detected in
zooplankton, benthic amphipod Hyallela Azteca, and fish species within 2 months of
addition (Harris et al.. 2007). indicating that Hg methylation and movement of MeHg
within the food web can occur within a single season in aquatic ecosystems. Estimated
residence time of atmospherically deposited Hg2+ in the water column of Little Rock
Lake, WI, was 150 days ("Watras et al.. 2006).
Methylmercury Cycling
Methylmercury (MeHg) is produced by sulfur-reducing prokaryotes (SRPs) in the course
of sulfate reduction. It is toxic to humans and vertebrate animals, particularly because it
biomagnifies up the food chain and bioaccumulates within organisms. The following
sections describe transfer of MeHg between ecosystem compartments, as well as
characteristics of zones of high MeHg production within the landscape.
Transfer of Methylmercury from Terrestrial to Aquatic Ecosystems
MeHg is a very small component (1%) of Hg atmospheric deposition to ecosystems
(Wcntz et al.. 2014); most MeHg in ecosystems is microbially produced from inorganic
Hg. Much of our knowledge about Hg cycling in natural ecosystems comes from the
Experimental Lakes Area (ELA) in northwestern Ontario, Canada, which receives low
Hg deposition [47,000,000 ng/ha/yr or 47 mg/ha/yr in 1990-1993; 80,000,000 ng/ha/yr
or 80 mg/ha/yr in 1998-1999 (St Louis et al.. 2001; St. Louis et al.. 1996)1. A small
portion of total Hg deposition in the ELA is in the form of MeHg, 900,000 ng/ha/yr
(reported as 0.9 mg/ha/yr, or <1.5% of total Hg deposition). MeHg cycles through the
terrestrial ecosystem as total Hg does. Leaf litter represents a significant flux of MeHg
from the living canopy in upland forest to the detrital food web of 800,000 ng
MeHg/ha/yr (reported as 0.8 mg MeHg/ha/yr), a flux of a similar magnitude to
atmospheric deposition (St Louis et al.. 2001). In the 20 years preceding the study,
600,000 ng MeHg/ha/yr (reported as 0.6 mg/ha/yr of MeHg) were annually stored in the
July 2018
12-105
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
organic soil pool, and on average 100,000 ng MeHg/ha/yr (reported as 0.1 mg
MeHg/ha/yr) was exported from the catchment in stream flow (St Louis et al.. 2001).
In five boreal forest catchments at the ELA, St. Louis et al. (1996) assumed a background
MeHg deposition rate of <500,000 ng MeHg/ha/yr (reported as <0.5 mg MeHg/ha/yr). A
catchment consisting only of upland forests exported 18% of estimated annual MeHg
deposition, indicating that terrestrial ecosystems retain MeHg (St. Louis et al.. 1996). and
that aquatic MeHg is produced endogenously.
Transfer of Methylmercury from Wetlands to Aquatic Ecosystems
At the landscape level, MeHg export from wetlands in streams and rivers is correlated
with wetland cover in watersheds. In the ELA in Ontario, a catchment with valley bottom
wetlands exported 73% of deposited MeHg in a dry year, but 158-200% of deposited
MeHg in a wet year, indicating that the wetlands produced MeHg in wet years. MeHg
export was high in catchments with other types of wetlands as well. A catchment with
riverine wetland (peatland fed by surface water flow) exported 120-130% of annually
deposited MeHg in dry and wet years, and a catchment containing a basin wetland
exported 213% MeHg in a dry year and 500-719% annually deposited MeHg in wet
years, indicating a strong source of MeHg within the watershed, presumably in the
wetlands (St. Louis et al.. 1996). More recent work in the watershed of Arbutus Lake, NY
measured Hg and MeHg in streams draining upland and wetland areas; MeHg was
1.7-3.0% of total Hg in upland streams and 7.8% of total Hg in a similarly low-order
stream that drained a wetland. In downstream peatlands, MeHg was 5.4-9.5% of total Hg
(Selvendiran et al.. 2008a). and the beaver meadow peatland exported 623% of total
MeHg input annually, while the riparian wetland exported 8% of total MeHg input
(Selvendiran et al.. 2008b). Wetlands with high dissolved organic carbon and fluctuating
water levels and anoxic zones are areas of high MeHg production within watersheds, and
such wetlands will affect MeHg concentrations in water and biota downstream.
Methylmercury Cycling in Aquatic Ecosystems—Lake Onondaga, New York
Lake Onondaga, NY was the subject of a Hg budget project in 1992 because the
eutrophic lake received high Hg loads through the late 1970s from industry
[Supplemental Figure 1; (Henry etal.. 1995)1.
July 2018
12-106
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
Net Methylalion
(0, 0.63)
Net Uptake by Fishes
(0.20, 0.20)
Gross
Sedimentation
(13.7, 0.73)
Dissolved Flux
(0.056, 0.017)
Remineralization (2.6, 0.25)
Net Sedimentation
(11.3, 0.62)
Tributaries Atmospheric
and Metro Deposition Volatilization Outlet
(13 6, 0,26) (0,44. 0.006) (0.016, 0) (2,8, 0.24)
^ Groundwater
Inflow
(0.02. 0.001)
Hg = mercury; kg = kilogram: MeHg = methylmercury; yr = year.
Source: from Henry et al. (1995).
Supplemental Figure 1 Mercury and methylmercury mass balance cycle in Lake
Onondaga in 1992. The quantities of mercury in each flux
are indicated in parentheses by (kg total Hg/yr, kg
MeHg/yr).
In 1992, .1.4.1 kg of total Hg entered the lake, 96% of which was carried into the lake by
tributaries, including one that drained a former chlor-alkali plant. The largest sink for Hg
within the lake was through sedimentation, which accounted for 11.1 kg of total Hg. In
that same year, 2.8 kg of total Hg flowed out of the lake downstream and 0.20 kg Hg
(0.1% of total annual Hg load) was incorporated into fish biomass (Henry et al.. 1995).
A budget for Hg methylation was also constructed for the lake, and differs from that of
total Hg in several important points (Figure 12-4). First, only 29% of the annual lake
MeHg load of 0.88 kg entered the lake via tributaries (compared to nearly the entire total
Hg load). The water column, particularly deep, anoxic water, contributed 0.60 kg or 68%
of the annual MeHg load of the lake (Henry etal.. 1995). More than half (about 0.47 kg)
of the annual MeHg load was incorporated into sediments, and while sedimentation rates
were high, about one- to two-thirds of MeHg in sediments was released back into the
water column. Because most of the total Hg in fish is in the form of MeHg, the model
calculated that the MeHg in fish biomass was the same amount as total Hg in fish
My 2018
12-107
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
biomass, 0.20 kg, although this value constituted a much higher portion of the annual
MeHg load [23% of annual MeHg; (Henry etal.. 1995)1. This early study showed that the
majority of MeHg in Lake Onondaga was produced within the lake itself, specifically in
the anoxic zone, and that MeHg cycled between the sediments and water column, with a
significant amount stored in fish tissue.
Lake Onondaga was resampled in 2006, 2007, and 2009, following industrial remediation
projects and incorporation of nitrification treatment in the municipal wastewater plant
discharging into the lake (Todorova et al.. 2014). In the 2006-2009 sampling, MeHg
concentrations in the epilimnion were significantly lower than in 1992, and total Hg in
the epilimnion was half of the total Hg concentrations measured in 1992 (Todorova et al..
2014). However, in all years there was a significant peak in MeHg concentrations and in
MeHg:tHG ratios in the epilimnion in the fall following turnover (Todorova et al.. 2014).
indicating that this seasonal event will have significant effects on total MeHg in the food
chain of the lake.
Methylmercury in Sediments and Water Column—Lakes
The sediments under shallow, oxygenated lake waters are an important site of Hg
methylation. Early literature showed that Hg methylation rates are higher in shallow
organic sediments than in deep clay-rich sediments in Southern Indian Lake in Manitoba,
Canada, and that the ratio of methylation:demethylation was also higher in organic
sediments (Ramlal et al.. 1986). In the Quabbin Reservoir, MA, MeHg fractions in
sediments were 3.1-16.3% at 1-m depth, and decreased with increasing lake depth to
0.1-0.2% at a 22 to 23-m depth (Gilmour et al.. 1992). At Lake Clara, WI, Hg
methylation rates were sampled in the water column, flocculant surface sediments, and
deeper sediments at water depths of 1-10.5 m (Korthals and Winfrey. 1987). At all
sampling depths, methylation and demethylation were detectable in the water column and
in the surface sediments. Gross Hg methylation rates were highest in surficial sediments
sampled at depths of 5-7.5 m, but the methylation:demthylation ratio was highest
(5.5-5.8) in surficial sediments at depths of 1-2 m (Korthals and Winfrey. 1987).
Larger, deeper lakes stratify during the summer; waters near the surface (epilimnion) of
the lake are high in oxygen and primary productivity, while deeper waters (hypolimnion)
have low oxygen, lower primary productivity, and high rates of S reduction and MeHg
production. In the experimentally acidified Little Rock Lake, WI, MeHg was low (less
than 0.1 ng/L) in surface waters during the summer months, but increased at anoxic
depths below 4-6 m to 1 ng/L in June, 3 ng/L in July, and 3.5 ng/L in August (Bloom et
al.. 1991). In stratified Lake Onondaga, NY, the water column, particularly deep, anoxic
July 2018
12-108
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
water, contributed 0.60 kg or 68% of the annual MeHg load of the lake (Henry et al..
1995). Sampling in Pallette Lake, WI, showed that the oxic-anoxic boundary was at 12 m
depth in this stratified lake, which was also the location of the sulfate-sulfide transition
(Watras et al.. 1995). There were peaks in both sulfate reduction and Hg methylation
from 12-13 m depth, right at the top of the hypolimnion, which was enriched in MeHg
(20-30%) compared to the epilimnion (5-10% MeHg). There was no sulfate reduction
detected in the profundal sediments or deep layers of the hypolimnion. MeHg was
transported from the oxic-anoxic boundary and hypolimnion where it was produced to the
epilimnion, and the flux of MeHg from the hypolimnion (3,400,000-6,800,000 ng/day,
reported as 3.4-6.8 mg/day) was much larger than the flux of MeHg into the epilimnion
from the atmosphere [10,500 ng MeHg/day, or 0.0105 mg MeHg/day; (Watras et al..
1995)1. Ecklev and Hintelmann (2006) showed that in boreal Canadian lakes, potential
methylation is highest approximately 1 m below the oxycline, and this methylation zone
moves up the water column as the oxycline rises due to oxygen depletion at depth in the
summer (Ecklev and Hintelmann. 2006).
Methylmercury in Sediments and Water Column—Wetlands
In wetlands at Bog Lake Fen in the Marcell Experimental Forest, MN, the highest levels
of MeHg production occurred at the upland-peatland interface. When mesocosms
installed in the bog interior received additions of labile C and sulfate, MeHg
concentrations were similar to MeHg concentrations from the upland-peatland interface,
indicating that upland inputs of C may contribute to high methylation rates in the
peatland (Mitchell et al.. 2008a). A larger study that sampled peatlands within the
Marcell Experimental Forest, MN, as well as at the ELA in Ontario found that %MeHg
was highest within 5-10 m of the upland-peatland interface across wetlands, and that
sulfate and pH were higher at the interface, and DOC was lower at the interface, than in
the interior of the peatlands (Mitchell et al.. 2008a). In peatlands in the Arbutus Lake
watershed in the Adirondack Mountains, NY, MeHg and total Hg were higher in pore
water near the top of the peat profile, at 20 to 40-cm depth, than at 80 to 100-cm depth
(Selvendiran et al.. 2008a). and the total Hg and MeHg were highest in both peat and
pore water in the top 15 cm of the peat (Selvendiran et al.. 2008a). The Hg in the top
15 cm of peat was estimated to be about 27% of the total Hg, and MeHg in the top 15 cm
was 30% of the total MeHg in a riparian peatland and 50% of the total MeHg in a beaver
meadow (Selvendiran et al.. 2008a). In freshwater Sphagnum peatlands of the Allequash
Creek watershed in the Northern Highland Lake District of Wisconsin, reducing
conditions exist below 2-cm depth in the peat, and MeHg concentrations peak at 7-cm
depth in the peat (Creswell et al.. 2008).
July 2018
12-109
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
In wetlands not dominated by Sphagnum, Hg methylation occurs in sediments and in
periphyton (see Appendix 12.3). In the freshwater marshes of the Florida Everglades,
where periphyton and flocculation often extend the sediment-water interface, methylation
rates were highest in the top 6 cm of sediments (Gilmour et al.. 1998). In coastal marsh
sediments, such as at Kirkpatrick Marsh on the Chesapeake Bay, MeHg was highest in
the top 6 cm of sediment, although total Hg was highest at a depth of 12-15 cm in the
sediments (Mitchell and Gilmour. 2008). and in coastal Georgia marshes, sediment core
incubations showed that MeHg production was highest in the top 4 cm of the marsh
sediments (King et al.. 1999).
Methylmercury in Sediments and Water Column—Estuarine and Marine
Ecosystems
Research shows that Hg methylation occurs in estuaries and in marine sediments.
Incubations of samples collected in the Patuxent River and Estuary, MD, showed that net
MeHg production in sediments was four times the rate of MeHg accumulation in
sediments, suggesting that MeHg in estuary surface waters comes from estuarine
sediment sources as well as sources upstream in the watershed (Benoit et al.. 1998).
MeHg comprised 5% of surface water particulate Hg (unclassified compounds large
enough to be removed by filtration) and 2% of total Hg in surface water, and 0.3% of Hg
in estuarine sediments (Benoit et al.. 1998). More recently, surface sediments were
sampled from the Chesapeake Bay, from four locations in the main channel, two
locations on the continental shelf, and one location on the slope of the continental shelf.
Total Hg in the upper bay ranged from 50 to 171 ng Hg/g sediment (reported as
250-850 pmol/g), in the lower bay and continental shelf ranged from 2.0 to 20 ng Hg/g
sediment (10-100 pmol/g), and at the single continental slope site was 50-70 ng Hg/g
sediment (250-350 pmol Hg/g); (Hollweg et al.. 2009). MeHg was 0.2-1.5% in the top
0-12 cm of sediments across sites, and bottom water salinity was strongly correlated with
%MeHg [r = 0.814; (Hollweg et al.. 2009)1. suggesting that MeHg was produced in
marine sediments.
An incubation study of water samples collected at depths of 20-327 m in the ocean
around the Canadian Arctic Archipelago found measurable rates of Hg methylation at all
depths from all five sites sampled. Methylation of inorganic Hg accounted for 47% of the
MeHg present in the water column at these locations, and demethylation was also widely
observed, indicating that marine MeHg represents marine production of MeHg, not
merely transport of MeHg from other systems (Lehnherr et al.. 2011).
July 2018
12-110
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
APPENDIX 13. CLIMATE MODIFICATION OF
ECOSYSTEM RESPONSE TO
NITROGEN AND SULFUR
The scope of this appendix is to identify key papers describing how climate alters
ecosystem response to nitrogen (N) and sulfur (S) addition. Nitrogen and S loading
occurs in many ecosystems concurrently experiencing multiple stressors, including
human-driven climate change. Climate change effects on U.S. ecosystems were recently
summarized in the U.S. National Climate Assessment (Galloway et al.. 2014; Groffman
ct al.. 2014). Each appendix of the ISA evaluating the effects of N enrichment or
acidification includes a section on how climate modifies the ecosystem response to N or
N + S deposition. Additionally, in this appendix, we have excerpted text from Greaver et
al. (2016). to serve as a foundation for the discussion. This paper, which focuses on
empirical observations, provides a current review of how climate (e.g., temperature and
precipitation) modifies ecosystem response to N.
Anthropogenic emissions of greenhouse gases are likely to cause a global average
temperature increase of 1.5 to 4.0°C and a significant shift in the amount and distribution
of precipitation by the end of the 21st century (Collins et al.. 2014). Recent work has
focused on the effects of anthropogenic N on the Earth's radiative forcing (Pinder et al..
2012) and how temperature and precipitation alter ecological responses to N exposure
(Greaver et al.. 2016). Most work is conducted on the effects of climate and N or
acidifying deposition (N + S). There are only a couple of studies on how climate modifies
ecosystem response to S.
Climate effects on ecosystems is a rapidly expanding field. For some processes we are
beginning to understand how temperature and precipitation may interact; however, for
many biogeochemical pool and processes, data are insufficient to quantify either the
direction or magnitude of how climate may alter ecosystem response to N with certainty.
Some global-scale earth systems models now incorporate interactions between N and
carbon (C) in ecosystems. They are summarized in Appendix 6. Greaver et al. (2016)
includes information on terrestrial and surface water ecosystems. A brief summary of
climate modification of estuary response to N is also included in Appendix 13.2.
June 2018
13-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
13.1. Climate Modification of Soil Acidification and Nitrogen
Enrichment
Climate-driven changes in N cycling may alter the N supply in terms of quantity and
timing of N available for uptake by biota. Alteration of N availability relative to a given
organism's life cycle or physiological thresholds may alter overall ecosystem function.
These effects may be further modified when temperature and precipitation cause direct
stress to biota. In the following sections we describe how climate alters the cycling of N
and how temperature and precipitation interact with two important mechanisms affected
by N availability to taxa: (1) nitrogen-driven eutrophication, which will stimulate the
growth of opportunistic plant and animal species, and (2) acidification, which may
decrease growth and cause mortality among sensitive species. We describe how these
changes in growth will alter C cycling and biodiversity.
13.1.1. Nitrogen Transport and Transformation
13.1.1.1. Excerpt from Greaver et al. (2016)
"Although global N cycling is complex, the movement of N through the biosphere can
largely be explained by describing a few key transformations (Figure 13-1). Atmospheric
N2 is converted into reactive N by lightning or by specialized bacteria capable of
biological N fixation, in addition to the human activities that create reactive N.
Organisms use reactive N to produce proteins and other essential compounds. Dead
organic matter is decomposed by microbial enzymes, producing smaller N containing
organic molecules such as amino acids. This organic N is largely converted to mineral
forms that are readily assimilated by plants and microorganisms. Where reactive N is
present under aerobic conditions, some microorganisms convert ammonium (NH/) to
nitrate (NO, ). a process termed nitrification. Nitrate is mobile in soils and often leaches
into aquatic systems and groundwater. In anaerobic conditions, microorganisms can
convert NO3 to gaseous N via denitrification, emitting N back to the atmosphere.
"In nonagricultural terrestrial ecosystems, atmospheric deposition is the dominant source
of anthropogenic N. Through changes in precipitation, shifts in atmospheric circulation,
and temperature-related effects on the stability of N compounds, climate change is
expected to alter the relative contribution of wet and dry forms of N deposition and shift
the spatial distribution of N deposition (Engardt and Langncr. 2013; Tagaris et al.. 2008).
Local N deposition rates are generally predicted to change by 0-20% (Engardt and
Langner. 2013; Tagaris et al.. 2008).
June 2018
13-2
DRAFT: Do Not Cite or Quote

-------
Spatial and temporal alteration of snow melt, precipitation and evapotranspiration
I
Spatial and temporal alteration of landscape-level wetness and hydrologic flow
i
Key mechanisms of nitrogen cycling altered due to changes in hydrology
Drier conditions
(-) Nitrogen flushing is likely to
increase N accumulation in
ecosystems, with large N pulses
exported during rainfall events.
(-) Dentrification will generally
decrease with drying, leading to
accumulation of N within the
ecosystem.
(-) Mineralization will generally
decrease under dry conditions.
(-) Nitrogen uptake by vegetation
caused by drought-stress in
vegetation as water becomes the
most limiting factor for growth.
(+) Drought-related plant infestation
and disease wilt generally decrease
under dry conditions.
(+) Fire will release ecosystem N into
the air and increase N available as
throughflow.
N = nitrogen.
Source Greaver et al. (2016).
Figure 13-1 Summary of key interactions between nitrogen,
anthropogenic-driven climate change, and hydrology.
1	"The influence of N addition on reactive N availability within terrestrial and aquatic
2	ecosystems is mediated by microbial transformations and transport within and between
3	ecosystems. Climate change is expected to strongly affect these processes through
4	increasing temperature and through temporal and spatial shifts in temperature and
5	precipitation (Baron et al.. 2012). Warming directly increases the metabolic biokinetics of
6	enzyme activity necessary for microbial N transformation until a temperature optima is
7	reached. Climate change is also expected to cause numerous modifications of the
June 2018	13-3	DRAFT: Do Not Cite or Quote
Wetter conditions
(+) Nitrogen flushing increases export
from 'upstream' ecosystem and
loading to the 'downstream'
ecosystem.
(+) Denitrification will generally occur
in wetter areas, increased
denitrification will cause more N in
the ecosystem to be lost via the
gas-phase.
(+) Mineralization will generally
increase under wet conditions.
(+) Moisture-related plant infestation
and disease will generally increase
under wet conditions.

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
hydrologic cycle, and moisture availability regulates the biokinetic temperature response
(Borken and Matzner. 2009; Rustad et al.. 2001) because water is needed to transport
enzymes and substrates and water influences oxygen availability. Warming is predicted
to intensify the hydrologic cycle, with heavy rainfall events expected to become more
frequent and intense, along with the potential for deepening and lengthening of dry
periods, as well as altered snow accumulation and melt and changes in evapotranspiration
(Collins et al.. 2014; Jimenez Cisneros et al.. 2014). These changes may cause soil
conditions for microbial activity to shift between optimal and inhibitory (Morse et al..
2015b). modifying the link between climate warming and the rate of microbial N
transformations such as decomposition, mineralization, nitrification, denitrification, and
biological N fixation. Of these transformations, the rates of N fixation may be the most
uncertain part of the N cycle (Vitousck et al.. 2013). altering N supply and influencing
the C cycle (Welter et al.. 2015).
"Climate-driven changes to the hydrologic cycle will also alter the quantity of N
transported through a system via waterborne transport and soil water content-mediated N
cycling (Billen et al.. 2013). Alteration of N retention in the soil due to changes in
moisture and flushing may be significant enough to determine whether an ecosystem is
an N source versus N sink (Boulton. 2007; Band et al.. 2001). Greater precipitation
generally increases water flow, which may (1) increase leaching/export of N through
terrestrial landscapes, (2) increase terrestrial N inputs to streams and rivers, and
(3) increase N transport rates through streams and rivers (Whitehead et al.. 2009).
although adaptation by microbes and plants may increase their ability to retain N as
flushing increases, thereby potentially limiting some of the overall impact of increased
precipitation and flow.
"Under dry conditions, landscapes can become more hydrologically disconnected and N
retentive, which can increase N concentrations in subsequent flushing events (Goodridgc
and Melack. 2012; Kaushal et al.. 2008). Drought can inhibit nitrification and cause N to
accumulate in the soil; once precipitation occurs it often results in a pulse of nitrification
that produces nitrate and subsequent nitrate leaching (Lamersdorf et al.. 1998). For
example, the 2012 droughts in the midwestern U.S. were followed during the spring 2013
by extremely high river nitrate concentrations (Beeman. 2013). Likewise, longer periods
between wet cycles lead to accumulation of nitrate and other acidifying solutes in the
soil, causing less frequent, yet more extreme acidification events (Bavlev et al.. 1992).
Beyond simply the total volume, precipitation intensity influences the rate of N flow
through ecosystems. Increased precipitation intensity of cold season frontal storm
systems and warm season convective storms would likely increase the frequency of high
N loading events to aquatic systems. Due to the limited capacity for in-stream removal of
13-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
N during high flow pulse events, most N is transported downstream (Kaushal et al..
2014V
"Such hydrologic cycle changes are also expected to affect the timing of N transport.
Changes in the seasonality of precipitation, and in particular snowmelt, will tend to alter
the timing of N flushing through the ecosystem. There has already been widespread
earlier snowmelt and increased winter thaws associated with warming over the last few
decades (Collins et al.. 2014). but implications for the timing of N export have been
assessed in only a few sites (Casson et al.. 2012). Timing changes ultimately can alter the
magnitude of N export to downstream water bodies, particularly if the timing of flushing
changes relative to the timing of biologically mediated uptake in either terrestrial or
aquatic ecosystems (Baron et al.. 2009). Thus, it is possible to have a modification in
sink/source behavior in regions where annual or seasonal patterns of water-filled pore
space shift with climate change.
"The rate at which denitrification returns reactive N to the atmosphere varies across space
and time, with landscape- to micro-scale denitrification hot spots or moments that depend
on interactions with hydrologic flow paths, the persistence and variability of low oxygen
conditions, and the residence time of water and N, all of which are likely to respond to
climate-driven changes in the hydrologic cycle (Anderson et al.. 2015; Weier et al..
1993). Generally, warmer and wetter conditions under climate change would facilitate
greater rates of denitrification, whereas warmer and drier areas might experience
decreased denitrification or concentrate "hot spots" into smaller areas with higher soil
moisture, substrate concentrations, and fluxes (Duncan et al.. 2013). Alternating wet and
dry states may promote coupled nitrification/denitrification processes or build-up and
flushing of mobile N depending on the ratio of transport to reaction rates. These are
general trends associated with moisture availability and transport. Carbon substrate
availability and other controls on microbial processes, however, will also be influenced
by climate change (discussed below) and can mediate these hydrologic effects."
13.1.1.2. Additional Considerations
There are considerations to note in addition to the excerpt from Greaver et al. (2016).
Additional papers on climate interactions with N and S soil biogeochemistry discussed in
Appendix 4 are summarized in Table 13-1. Featured most notably are the effects of
precipitation on S cycling, snow on soil N cycling, N addition effects on biogenic GHG
flux, and the results of integrating climate parameters into N and S biogeochemistry
modeling.
June 2018
13-5
DRAFT: Do Not Cite or Quote

-------
Table 13-1 Summary on climate modification sulfur and nitrogen cycling in
Appendix 4 and Appendix 6 in addition to those in Appendix 13.
Indicator/Process T P Snow
Reference
Biogeochemical effects (Appendix 4)
S mineralization X X	Empirical: Watershed mass balances of sulfur are
increasingly regulated by watershed moisture
because high-moisture soil conditions stimulate the
net mineralization of soil sulfur pools.
Mitchell etal. (2011)
S retention
X	Model: Watersheds with higher runoff ratio tend to
convert sooner from net retention to net release of
SCM2".
Rice etal. (2014)
Soil N variables
X Meta-analysis: Meta-analysis data is from
41 publications based on snow depth manipulation
experiments. The general responses of 12 variables
related to terrestrial nitrogen (N) pools and dynamics
to altered snowpack depth are evaluated.
Li etal. (2016c)
N mineralization X
X Empirical: More freeze/thaw cycles anticipated with
climate warming; supported lower rates of N
mineralization.
Duran et al. (2016)
N mineralization X
X Empirical: Snow-manipulation experiment at
Hubbard Brook to evaluate soil freezing later in the
season due to prolonged warm season. Freezing
likely increases soil NO3" levels by physical disruption
(increased fine root mortality) causing reduced N
uptake by plants and reduced competition for
inorganic N, allowing soil NO3" levels to increase
even with no increase in net mineralization or
nitrification.
Groffman et al.
(2001)
N mineralization
X Empirical: N mineralization rates were more strongly Sorensen et al.
related to soil volumetric water content than to root (2016)
biomass, snow or soil frost, or winter soil
temperature.
N effects CO2, CH4,
and N2O
Meta-analysis: 313 observations across 109 studies Liu and Greaver
evaluated the effect of N addition on the flux of three (2009)
major GHGs: CO2, CH4, and N2O.
Soil acidification and X X	Model: 2009-2100—six climate change simulations Pourmokhtarian et
N enrichment	of temperature, precipitation, and photosynthesis al. (2012)
active radiation (PAR). Without CO2 fertilization
effects, net soil mineralization and nitrification
increased, soil and stream acidified, and the
percentage of base saturation in soils declined due to
increased NO3". With CO2 fertilization effects, the N
loss to streams was suppressed due to increased
plant uptake.
June 2018
13-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
Table 13-1 (Continued): Summary on climate modification of nitrogen (N) effects
on nitrogen cycling in Appendix 4 and Appendix 6 in
addition to those in Appendix 13.
Indicator/Process T P Snow	Reference
Soil acidification, N XX	Model: The ForSAFE-VEG model simulations show Belvazid et al
enrichment, and	that climate and atmospheric deposition have	(2011a)
plant community	comparably important effects on N mobilization in the
soil, as climate triggers the release of organically
bound nitrogen stored in the soil during the elevated
deposition period. Climate has the most important
effect on plant community composition; thus climate
change cannot be ignored in future simulations of
vegetation dynamics.
Soil acidification and X X	Model: PnET-BGC modeling found climate change Wu and Driscoll
N enrichment	played a larger role in ANC than base cation	(2010)
deposition changes. Temperature had a larger effect
than precipitation on decomposition. Net
mineralization and nitrification increased faster with
climate change than plant NO3" uptake.
13.1.2. Nitrogen, Climate, and Carbon Cycling
13.1.2.1. Excerpt from Greaver et al. (2016)
"Changes in N supply alter plant growth and C cycling. The addition of N to terrestrial
and aquatic ecosystems can cause eutrophication, a state of high nutrient availability that
alters ecosystem function. Autotrophs (plant/algae) capture CO2 through photosynthesis,
storing C in biomass until it is oxidized through respiration or combustion and released
back to the atmosphere. In terrestrial systems, N addition usually stimulates autotrophic
growth until biotic N demand has been satisfied, although high rates of N addition may
increase the concentration of acid anions, which often decreases plant growth (see
acidification discussion below). At certain sites, N additions have uneven effects,
stimulating growth of some tree species, while impairing the health and growth of others
(Thomas et al.. 2010). When considering the net effects on multiple tree species, growth
in most forests is stimulated. This additional growth increases the overall amount of C
stored in plant biomass; one unit of N input may cause an additional 24.5 to 177 units of
forest C uptake (dc Vries et al.. 2014a; Liu and Greaver. 2009).
"A number of published meta-analyses evaluate the response of C pools and fluxes to
single stressors of N, precipitation, or temperature. To gain insight on stressor
interactions from single stressor response studies, we have synthesized existing
meta-analyses of terrestrial ecosystem response to additions ofN, precipitation, and
June 2018
13-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
temperature (e.g., [see Figure 13-2 in this document]). A recent correlation analysis of
growth (in terms of net primary production |NPP|) for 1,247 woody plant communities
across global climate gradients confirms that NPP increases with higher temperature and
precipitation (Chu et al.. 2016). However, ecological changes along broad natural
gradients may differ from the response of ecosystems to comparatively rapid
environmental change caused by human activities (Chu et al.. 2016). This latter process
may be more accurately characterized by addition experiments, such as those summarized
by meta-analysis. Our synthesis of existing meta-analyses indicates that aboveground
NPP is highly responsive to N addition and enhanced precipitation, while temperature
rise does not increase aboveground NPP. This result is consistent with the basic
biokinetic effects of warming on enzyme activity, which would have the counteracting
effects of stimulating both plant C capture (photosynthesis) and plant C release
(autotrophic respiration). Although there are no meta-analyses on warming and
whole-ecosystem autotrophic respiration, Lu et al. (2013 i observed in a meta-analysis
that temperature increased gross ecosystem production, a metric that does not subtract
respiration from gross production (photosynthesis). Therefore, temperature may have
larger effects on plant C fluxes than on NPP. Note that precipitation-induced changes in
NPP may vary depending on whether there is sufficient enhancement of precipitation to
offset increased evapotranspiration in a wanner climate (Bachclct et al.. 2001).
Carbon pools
Ecosystem C
31 ? ?
60
30
0
-30
N T P
GEP

NEE


R
? 34 48

31 34
48

? 34 48

60-

—i—
60

mR
30-
NS NS

30-



i—r	




-30-1


-30-1

Above-ground biomass
37 34 48
60-
30
0
-30
Foliage biomass
34 ? ?
60-
30-
0-
-30
Root biomass
37 34 48
60-
30-
0
-30
a
Fine root biomass
? 99 ?
60-
NS

N T P
60
30-
0
-30
ŁL
N T P
60
30-
0
-30
NS
Above-ground NPP Below-ground NPP Below-ground RautaJrC(lf1s
97 48 48	? 48 48	? 34 100
60-
30-
0
-30
N T P
Organic layer
36 34 ?
Soil C Dissolved organic C Microbial C	Myrcorrhiza
36 34 ?	36 34 ?	36 34 ?	38 ? ?
Decomposition
98 34 ?
60-
30-
0-
-30-
1 NS
EHi—i
60-
30-
0
-30
H2	
N T P
60-

60
30-

30

NS,		

-30-

-30
36 34 100
h«
-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
Note: Bars indicate response ratios (treatment/control x 100), and error bars represent 95% confidence intervals for the response.
Orange bars indicate the magnitude of response to nitrogen enrichment, blue bars show response to temperature increase, white
bars show response to precipitation increase. Blue zone indicates ecosystem carbon inputs and outputs, green zone indicates plant
responses, and brown zone indicates soil and microbial responses. The upper confidence interval for the precipitation effect on net
ecosystem C exchange is 124.6 and beyond the scale of the chart. The response of aboveground net primary production to
warming was stated to be nonsignificant (Wu et al.. 2011). but no effect size was given.
Source: Greaver et al. (2016) The number above each bar indicates the published source of the effect size as follows: 1. LeBauer
and Treseder (2008): 2. Wu et al. (2011): 3. Lu et al. (2013): 4. Liu and Greaver (2009): 5. Liu and Greaver (2010): 6. Xia and Wan
(2008): 7. Knorr et al. (2005): 8. Dieleman et al. (2012): 9. Treseder (2004): 10. Luetal. (2011b): 11. Liu et al. (2016b).
Figure 13-2 The effects of increased nitrogen, temperature, and precipitation
on terrestrial carbon pools (left panel) and fluxes (right panel)
from published meta-analyses.
"Belowground, initial findings are that N addition tends to increase the C stored in the
soil organic layer and in root biomass (Liu and Greaver. 2010; Xia and Wan. 2008).
while it tends to decrease mycorrhizae/microbial abundance and heterotrophic respiration
(Liu and Greaver. 2010; Treseder. 2004). This offset may result in no net change in soil
respiration (Liu and Greaver. 2010); however, this is an active area of research.
Consistent with the biokinetic effects of warming, long-term data and meta-analyses
show that soil respiration, including decomposition and microbial respiration, is
stimulated by increasing temperature (Lu et al.. 2013; Bond-Lambertv and Thomson.
2010; Rustad et al.. 2001). Most empirical studies show rising temperature stimulates N
release by mineralization (Churkina et al.. 2010). which may be driven more by
temperature effects on moisture (Emmett et al.. 2004). In some dynamic land models, the
additional N from mineralization will stimulate C uptake by plants even more than
current N deposition (Burd et al.. 2016). At the same time, increased N from
mineralization may cause N induced inhibition of decomposition, a feedback mechanism
that might decrease the amount of N released that is currently considered by few models
(Gerber et al.. 2010). The mechanisms causing N driven reduction in decomposition are
not well understood, but are thought to result from changes in microbial community
composition and their production of decomposition enzymes, as well as possible changes
in the character and degradability of soil organic matter (Conant et al.. 2011; Janssens et
al.. 2010). Climate change could also affect decomposition rates by altering both
available soil moisture and microscale connectivity among microorganisms, water, and
nutrients within the soil matrix that in turn may alter microbial processes (Xiang et al..
2008). Although there is no consensus about how dissolved organic carbon (DOC) in
surface water is regulated overall, increasing N and temperature increase DOC
concentrations (Laudon et al.. 2012). While few meta-analyses examine precipitation
effects on the soil C cycle, precipitation tends to increase the root C pool (e.g., [see
Figure 13-2 in this document])
June 2018
13-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
"Overall ecosystem C balance is assessed by summing measurements of individual pools
to quantify ecosystem carbon (EC) content or by measuring C fluxes to quantify the net
ecosystem exchange (NEE) of C. The meta-analysis we identified indicated that
temperature did not increase NEE lYLu et al.. 2013) positive NEE indicates ecosystem C
gain], and as previously mentioned this is likely because the biokinetic effects of
warming stimulate respiration that offsets the C capture via stimulation of primary
production. Increased precipitation tends to increase NEE (Wuetal.. 2011). likely
because water availability increases photosynthesis, while not increasing plant respiratory
losses. A meta-analysis of N addition studies indicates that adding N to grasslands had no
effect on NEE, but that N addition increased forests EC (Liu and Greaver. 2009). There
may be differences between grasslands and forests in terms of the extent that C gain
simulated by N is offset by heterotrophic and autotrophic respiration. Increasing
temperature may decrease C storage if the warming causes inhibition of photosystems, or
enhances evaporation and reduces water availability. A better understanding of these and
other contributing processes is needed.
"Traditionally, primary production in freshwater systems was thought to be phosphorous
(P) limited, but recent data have shown an increase of limitation by N or colimitation by
N and P (Elser et al.. 2007). In N limited fresh waters, N addition enhances rapid growth
of nitrophilic algae. This is an important food source to consumer species in the trophic
cascade; however, it is unclear if this is an important source of long-term C storage.
Sediments are estimated to be the largest pool of long-term C storage (Cole et al.. 2007).
and N stimulates the production of terrestrial biomass that may be transported to aquatic
sediments. Nitrogen also stimulates primary production of aquatic algae which contribute
to C in sediments (Kastowski et al.. 2011). although few studies have examined this
effect. Gudasz et al. (2010) found a strong positive relationship between increasing
temperature and organic C mineralization. They conclude future organic C burial in
boreal lakes could decrease 4-27% under scenarios of warming due to enhanced
temperature-dependent microbial activities. We were unable to identify studies
examining precipitation effects, or the combined effects of N, temperature, and
precipitation effects, on C storage in freshwater ecosystems. A discussion of biodiversity
associated with eutrophication of fresh waters is included in the biodiversity section."
13.1.2.2. Additional Considerations
There are considerations to note in addition to the excerpt from Greaver et al. (2016).
Additional papers on the interacting effects of climate and N on C cycling discussed in
Appendix 4 and Appendix 6 are summarized in Table 13-2. Notably there is a new
June 2018
13-10
DRAFT: Do Not Cite or Quote

-------
meta-analysis on N and warming interactions and several new papers on the interacting
effects of N on plant biomass and precipitation.
Table 13-2 Summary on climate modification of nitrogen (N) effects on carbon
(C) cycling in Appendix 4 and Appendix 6 in addition to those in
Appendix 13.
Indicator/Process T P Snow	Reference
Biogeochemical effects (Appendix 4)
19 different C pools X X	Synthesis: A synthesis of meta-analyses for single Greaver et al. (2016)
and processes	factor experiments on N, T, or P manipulation for
within ecosystems	19 different C pools and processes within
ecosystems.
Soil C	X X	Meta-analysis: Results showed that the interaction of Ni et al. (2017)
warming and N deposition greatly increased the soil
C input (+49%) compared with the single factor of
either warming (+5%) or N deposition (+20%). Soil C
loss was not significantly affected by the interaction of
N and warming, likely because increases in
decomposition due to warming are offset by the
decreases by N addition.
N effects CO2, CH4,
and N2O
Meta-analysis: 313 observations across 109 studies
evaluated the effect of N addition on the flux of three
major GHGs: CO2, CH4, and N2O.
Liu and Greaver
(2009)
Decomposition X
Empirical: Increasing rainfall variability and N
addition can stimulate litter decomposition in tall
grass prairies in the U.S.
Schuster (2016)
Bioloqical effects (Appendix 6)
Biomass X
Empirical: Using 1,600 observations, the study
authors found that biomass responses to N increased
linearly with mean annual precipitation (MAP).
Xia and Wan (2008)
Biomass X
Empirical: Using 126 observations, the study authors
found no relationship between biomass and mean
annual precipitation (MAP).
LeBauer and
Treseder (2008)
ANPP	X X	Empirical: Plant growth response to N increased with Tian et al. (2016a)
precipitation until annual precipitation reached
800 mm/yr. The N response efficiency also peaked
with moderate annual temperatures (~8°C) and
declined under cooler or warmer conditions.
June 2018
13-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Table 13-2 (Continued): Summary on climate modification of nitrogen (N) effects
on carbon (C) cycling in Appendix 4 and Appendix 6 in
addition to those in Appendix 13.
Indicator/Process T P Snow	Reference
ANPP	X	Empirical: In the Sonoran Desert near Phoenix, AZ, Hall et al. (2011)
60 kg N/ha/yr caused little or no increase in
production among herbaceous annuals in low
precipitation years, moderate N responses with
average rainfall, and strong increases in biomass with
added N during above-normal rainfall seasons.
Aboveground	X	Empirical: Added N was positively correlated with Vourlitis (2012)
biomass and litter	precipitation and was only significant in the high
production	rainfall years.
13.1.3. Climate and Acidification
13.1.3.1. Excerpt from Greaver et al. (2016)
"Atmospheric deposition of acidic N and sulfur (S) compounds (e.g., HNO3, H2SO4) has
directly caused widespread acidification of terrestrial and freshwater ecosystems in many
industrial areas. More broadly, other forms of atmospheric N deposition (e.g., NHx) can
indirectly cause acidification by increasing inorganic N availability enough to induce
nitrification. Freshwater and terrestrial ecosystem acidification is well studied and is
characterized by decreased pH and elevated aluminum (Al) concentrations/mobility in
soils and surface waters that cause plant physiological changes (Schaberg et al.. 2002).
tree mortality (McNultv et al.. 2007). aquatic fauna mortality, and decreased aquatic
biodiversity (Driscoll et al.. 2001b). Acidification driven by N (as opposed to N + S)
occurs at higher levels of N addition than for initial changes to the C cycle. Often N
saturation of the terrestrial ecosystem and subsequent leaching into adjacent aquatic
systems is observed in the process of aquatic acidification (Aberetal.. 1998). The
threshold for the onset of acidification changes across the landscape, depending on
geochemical sensitivity and historical loading of acidifying deposition. Although recent
declines in emissions of SOx and NOx in eastern North America and throughout Europe
(NAPAP. 2011) have led to many improvements in acid-base balances in acid sensitive
ecosystems in recent decades, it is unclear whether sensitive ecosystems will continue to
improve as emissions decline or whether secondary processes will promote, arrest, or
reverse ecosystem recovery as climate changes.
"Potential future changes in the quantity and temporal distribution of precipitation and
temperature (and their interactions) is expected to alter the wet-dry cycles that govern the
June 2018
13-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
timing and amount of acidic inputs in precipitation, microbial transformation in the soil,
and the flush of acid anions from soils to surface waters. If acid anions build up in soil
during periods of drought, the eventual flushing likely causes a more potent acidification
event (Moslcv. 2015; Whitehead et al.. 2009). If the acidification event occurs during a
time when sensitive biota or lifestages of biota are present, acidification may cause more
adversity to these populations (Kowalik et al.. 2007). Increases in storm frequency
associated with global climate change (Collins et al.. 2014) could increase the frequency
and severity of acidification driven by high levels of sea salt deposition in coastal regions
(Wright and Schindler. 1995). Although the mechanisms of interaction are unclear,
increases in DOC concentrations in aquatic ecosystems across Europe and the U.S. have
been linked to acidification, N cycling, and climate change, with important implications
for water quality and ecosystem function (Evans et al.. 2008).
"As previously mentioned, warmer temperatures increase decomposition and
nitrification. Nitrification will also increase with increased N supply caused by increased
weathering or decomposition (Booth et al.. 2005). The process of nitrification generates
protons that increase the rate of nitrate and base cation leaching to drainage waters
(Murdoch et al.. 1998). The combined increase of NO3 leaching and loss of base cations
has the potential to magnify acidification in forest soils (Fernandez et al.. 2003). Soil
weathering is typically the key buffer to acidic deposition (Li and McNultv. 2007). and
while weathering is increased by both soil temperature and soil moisture (Gwiazda and
Broecker. 1994). it is unclear whether any future change in the magnitude of temperature
and precipitation will be enough to alter base cation supply or influence the acid-base
balance of sensitive ecosystems. Furthermore, it is unclear whether increased supply of N
in soils from either deposition, increased decomposition, or increased nitrogen fixation
may negate the ameliorative effect of enhanced weathering. Some studies show that
climate change will mitigate acidification through increased weathering (Belvazid et al..
2011a). while others show that climate change will aggravate acidification although
increased nitrification outpacing enhanced weathering (Wu and Driscoll. 2010). In
general, increased temperature and precipitation will likely enhance inputs of buffering
agents from weathering and deposition, but also increase inputs of acidifying agents from
deposition and enhanced N cycling. The relative sensitivity of these opposing processes
to a given change in climate remains unresolved.
"Climate change may alter the sensitivity of biota to acidification, creating the need to
adapt to a combination of acidity and climate change stresses. For example, the suitable
habitat for brook trout in the Catskills and Adirondack mountains of the northeastern U.S.
may be constrained as climate change increases downstream water temperatures,
reducing downstream range where the trout can survive, while upstream migration is
limited in part by acid conditions in the headwaters. In another example, Al is toxic to
13-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
many fish and known to be mobilized during acidification events. It is known that the
mortality rate of Atlantic salmon exposed to A1 increases at higher temperatures (ABS
and Muniz. 1993); it is unclear how many other aquatic species would experience
temperature-dependent toxicity, which could make them more vulnerable to acidification
in a warming climate. Overall, there is little knowledge of how the biological thresholds
to acidity will be affected by climate change."
13.1.3.2. Additional Considerations
There are considerations to note in addition to the excerpt from Greaver et al. (2016).
Additional papers on climate interactions with acidification discussed in Appendix 4 and
Appendix 5 are summarized in Table 13-3. Most notably, several new studies model the
interacting effects of climate and deposition on soil biogeochemistry, plant growth, and
plant community composition.
June 2018
13-14
DRAFT: Do Not Cite or Quote

-------
Table 13-3 Summary on climate modification of acidification in Appendix 4 and
Appendix 5 in addition to those in Appendix 13.
Indicator/Process T P Snow	Reference
Biogeochemical effects (Appendix 4)
Soil acidification X X	Model: 2009-2100—six climate change simulations Pourmokhtarian et
and N enrichment	of temperature, precipitation, and photosynthetically al. (2012)
active radiation (PAR). Without CO2 fertilization
effects, net soil mineralization and nitrification
increased; soil and stream acidified; and percentage
of base saturation in soils declined due to increased
NO3". With CO2 fertilization effects, N loss to streams
is suppressed due to increased plant uptake.
Model: The ForSAFE-VEG model simulations show Belvazid et al.
that climate and atmospheric deposition have	(2011a)
comparably important effects on N mobilization in the
soil, because climate triggers the release of
organically bound nitrogen stored in the soil during
the elevated deposition period. Climate has the most
important effect on plant community composition,
underlining the fact that this cannot be ignored in
future simulations of vegetation dynamics.
Soil acidification X X	Model: PnET-BGC modeling found climate change Wu and Driscoll
and N enrichment	played a larger role in ANC than base cation	(2010)
deposition changes. Temperature had a larger effect
than precipitation on decomposition. Net
mineralization and nitrification increased faster with
climate change than plant NO3" uptake.
Biological effects (Appendix 5)
Growth of red	X X	Empirical: Tree ring analysis was conducted along Wason et al. (2017)
spruce and balsam	an elevation gradient (proxy for climate change). Both
fir	species showed increased growth with increased
precipitation pH. Red spruce appeared to show
increased growth with a warming climate but balsam
fir did not. No changes were noted in the [species?]
distribution in the spruce-fir forest, and the authors
suggest any such change may be part of a longer
term process.
Model: The Annual Radial Model used to investigate Koo et al. (2014)
projected climate change effects and changing air
pollution on red spruce growth. High-elevation
(<1,700 m) red spruce growth would decline if climate
change co-occurred with an increase in air pollution.
However, growth would increase under climate
change, if air pollution decreased. In contrast,
low-elevation red spruce growth would decrease with
future climate change regardless of change in air
pollution.
Soil acidification, N X X
enrichment, and
plant community
Growth of red	X X
spruce
June 2018
13-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Table 13-3 (Continued): Summary on climate modification of acidification in
Appendix 4 and Appendix 5 in addition to those in
Appendix 13.
Indicator/Process T P Snow	Reference
Biological effects of freshwater acidification (Appendix 8)
Water DOC	X	The major stressor affecting native coldwater fish Warren et al. (2017)
species in the eastern U.S. is shifting from
acidification to thermal stress and some lakes
recovering from acidification may provide a degree of
protection against climate effects. As DOC in the
water increases with increasing lake pH in recovering
lakes, decreased water clarity may create cooler
refuge habitat for fish.
13.1.4. Nitrogen, Climate, and Biodiversity
13.1.4.1. Excerpt from Greaver et al. (2016)
"Biodiversity, which contributes to the structure and function of ecosystems, is declining
globally (Rockstrom et al.. 2009). Decades of study show that added N reduces
autotrophic diversity in terrestrial and aquatic systems, fungal biodiversity in soils, and,
although less studied, animal diversity in terrestrial and aquatic systems (Howarth et al..
2011; Bobbink et al.. 2010). As previously mentioned, two mechanisms that contribute to
altered biodiversity are eutrophication and acidification. Eutrophication often causes N
stimulated growth for opportunistic species, which may cause competitive exclusion of
poorer competitors and soil acidification driving cation imbalances and physiological
stresses, suppressing seed germination and seedling regeneration (Sullivan et al.. 2013;
Roem et al.. 2002). In addition, N may alter physiology and/or community properties,
increasing the risk to secondary factors such as pests, fire, frost, and drought (Bobbink et
al.. 2010). Lastly, direct damage to vegetation from ammonia (NH3), nitrogen dioxide
(NO2), nitrogen monoxide (NO), peroxyacetyl nitrate (PAN), and nitric acid (HNO3)
exposure is known to occur, but most likely occurs in highly polluted areas and in close
proximity to high emission sources (Greaver et al.. 2012). In terrestrial ecosystems, all
processes generally reduce local autotroph diversity and homogenize habitats into
communities with small numbers of generally fast-growing or acid-tolerant autotrophic
species. These changes propagate through the food web, leading to increases of generalist
pests, herbivores, and parasitic soil bacteria, and decreases in specialist herbivores and
beneficial microbial communities that occur belowground (de Sassi et al.. 2012).
June 2018
13-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
"Biodiversity responses on land may be moderated by the type of climate change
occurring and the mechanism of N response. In terrestrial systems that get warmer and
wetter, eutrophication may be amplified if endogenous N sources are low or dampened if
endogenous sources are high/more liberated to meet community demand. The response of
the acidification pathway may depend on whether the change in net fluxes of cations
from climate change exceeds the net fluxes of N [i.e., from enhanced deposition of
cations and N, decomposition or weathering, and leaching (Howarth et al.. 2011; Roemet
al.. 2002)1. Many of these processes may be dampened in terrestrial systems that are
anticipated to get warmer and drier (e.g., southwestern U.S.) due to the drier conditions
reducing biological activity (Bobbink et al.. 2010). Colder regions such as montane,
alpine, and tundra systems are often strongly N limited and poorly buffered; thus, an
extended growing season under climate change will lead to greater opportunities for all
operating processes. Climate change may also magnify the effects of secondary stressors
in several ways, including increased pest populations under warmer wetter conditions, as
well as increasing fire potential and drought vulnerability as more aboveground plant
tissue is produced under elevated N (Pise et al.. 2011). Nonetheless, field evidence for
interactions between N and climate change under controlled conditions is scarce. The few
existing studies find additive effects in Mediterranean California (Zavalcta et al.. 2003).
and no interactive effects of precipitation and N addition in Minnesota (Reich et al..
2014; Reich. 2009). Not all terrestrial ecosystems are anticipated to be equally sensitive
to these pressures. In grasslands, many forbs and slow-growing species such as native C4
grasses appear especially vulnerable to added N (Clark and Tilman. 2008; Zavaleta et al..
2003). In a study of northeastern forests, all three tree species with negative growth
responses to N deposition were evergreen conifers (e.g., Pinus resinosa, Picea rubens,
Thuja occidentalis), while all five tree species with positive growth responses were
broadleaf species with arbuscular mycorrhizal associations [e.g., Acer rubrum, A.
saccharum, Fraxinus americana, Liriodendron tulipifera, and Prunus serotina (Thomas
et al.. 2010)1. However, contingent factors underly these general patterns, as there were
tree species from each group that did follow these generalities.
"In aquatic systems, elevated temperatures and N inputs from increased rain and glacial
retreat will likely magnify changes in algal assemblages that can propagate through the
food web (Hobbs et al.. 2010; Elseretal.. 2009a). In freshwater aquatic biodiversity
research, there is a substantial amount of work on lakes investigating the effects of
warming via gradient studies (latitude or altitude), warming experiments, time series, and
palaeoecology (Jeppesen et al.. 2014). Fish community assemblages, size, structure, and
dynamics are likely to change with continued global warming, and in some cases the
elevated temperatures that have already occurred in the past decades. Fish cannot
thermoregulate, but only physically move to areas with appropriate temperatures, if those
are accessible. In general, changes in fish composition, particularly in shallow lakes, are
13-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
characterized by a decline in abundance of several cold-stenothermal species (Kangur et
al.. 2013; Winfield et al.. 2010) and in increase in eurythermal species, which exhibit a
wide range of thermal tolerance (Jcppcscn et al.. 2012). Many fish species are also
adapted to specific oxygen concentrations; when temperature increases, oxygen may drop
to critical levels as warm water holds less oxygen and the respiration rates increase.
Warming effects on the biodiversity of grazing macroinvertebrates and zooplankton is
mixed across studies (Qzcn et al.. 2013; Burgmer et al.. 2007). Warming is shown to
increase cyanobacteria biomass (Kostcn et al.. 2012) and biofilm biomass (Williamson et
al.. 2016). while warming effects on phytoplankton show mixed results (Mccrhoff et al..
2012; Feuchtmavr et al.. 2009).
"Warming may cause increases in evaporation that will lower water level and increase
salinity ("Williams. 2001). Increasing salinity of freshwater systems tends to have a
negative effect on phytoplankton, zooplankton, macroinvertebrates, and fish (Jeppesen et
al.. 2007). Climate change will alter the transport, availability, and timing of N in
ecosystems, and recent data indicates an increase of primary production limitation by N
or colimitation by N and P (Elser et al.. 2009a). In N-limited fresh waters, N enrichment
enhances rapid growth of nitrophilic algae that outcompete other populations for light and
resources such as P and silicon (Si), leading to dominance by a few algal species and a
reduction in the nutritional quality of invertebrates as food for fish (Baron et al.. 2012;
Hobbs et al.. 2010; Elser et al.. 2009a). The combination of increasing both N and
temperature may be synergistic and sometimes difficult to uncouple (Kangur et al..
2013). as both may stimulate hypoxic conditions, consequently altering community
structure (Ozen et al.. 2013; Feuchtmavr et al.. 2009) and the frequency, intensity, extent,
and duration of harmful algal blooms (Paerl etal.. 2011).
"Presently, there are at least 78 listed or candidate species for threatened or endangered
status in North America that have N impacts identified as a primary contributor, and an
estimated 15-37% of species may be at risk from climate change (Hernandez et al..
2016). In total, there are numerous pathways whereby these dominant global change
factors can interact to impact biodiversity, and it is likely, although not definitive, that N
and climate often have additive and potentially amplifying effects on decreasing
biodiversity in many systems."
13.1.4.2. Additional Considerations
There are considerations to note in addition to the excerpt from Greaver et al. (2016).
Additional papers on climate interactions with N and biodiversity discussed in
Appendix 6 are summarized in Table 13-4. Most notably there are several new studies on
June 2018
13-18
DRAFT: Do Not Cite or Quote

-------
1
2
the interacting effects of climate and N deposition on plant species richness, and
community composition of plants, soil microbes and soil animals.
Table 13-4 Summary on climate modification of nitrogen (N) effects on
biodiversity in Appendix 6 in addition to those in Appendix 13.
Indicator/Process
T
p
Snow
Reference
Plant species
richness
X

Empirical: No significant interaction between N and
temperature on plant species richness for
closed-canopy systems (deciduous, evergreen, and
mixed forests), but significant interaction in
open-canopy systems (grasslands, shrublands, and
woodlands). In these open-canopy ecosystems, N
had a more negative effect on species richness at
lower temperatures.
Simkin et al. (2016)
Plant community
composition
X
X
Model: In the ForSAFE-VEG model projections of
plant community composition in three French forests
two N reduction scenarios showed that recovery of
these plant communities occurred only if climatic
factors are held constant at current levels.
Rizzetto et al. (2016)
Plant community
composition
X
X
Model: Understory plant community composition in
northern hardwood forests at Bear Brook Watershed
in Maine and Hubbard Brook in New Hampshire: the
simulated plant community composition returned
toward preindustrial conditions over the next century
only in a scenario in which N deposition rates
returned to background and climate was kept stable.
Phelan et al. (2016)
Soil microbial
community

X
Empirical: N additions and lower precipitation
interacted to shift microbial community composition in
forests soils in northeastern China.
Wanq et al. (2014c)
Soil microbial
community

X
Empirical: N additions and lower precipitation
interacted to shift microbial community composition in
grassland soils in Inner Mongolia.
Li et al. (2016a)
Ectomycorrhizal
species
composition

X
Empirical: Both changes in precipitation and nitrogen
were associated with shifts in ectomycorrhizal
species composition in European Scots pine.
Jarvis et al. (2013)
Nematode
predators,
microarthropod
herbivores, and
taxa richness of
nematodes and
microarthropods

X
Empirical: In Minnesota grassland, a decline was
observed in several categories of biota within a soil
food web (numbers of nematode predators,
microarthropod herbivores, and taxa richness of
nematodes and microarthropods) under increased N,
and a decrease in ciliate protists under high N and
drought.
Eisenhauer et al.
(2012)
June 2018
13-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
Table 13-4 (Continued): Summary of climate modification of nitrogen (N) effects
on biodiversity in Appendix 6 in addition to those in
Appendix 13.
Indicator/Process T P Snow	Reference
Algal species	X	In regions with no evidence of increased atmospheric Ruhland et al
composition	nutrient inputs, warming trends are observed to	(2015).
enhance competitiveness of planktonic diatoms like
Asterionella formosa, which are typically associated
with elevated N, indicating that climate change has
significant direct and indirect effects on algal species
composition. Climate change may thus enhance the
effects observed in areas with nutrient increases
alone.
13.2. Estuaries
In addition to terrestrial and freshwater systems described above, climate change
modifies key processes in estuarine and near-coastal systems linked to nutrient inputs.
Atmospheric deposition may represent more than 40% of N in coastal water bodies
(Appendix 7). N sources to near-coastal ecosystems include direct deposition to the water
surface and all other N sources upstream. Coastal eutrophication is a process of
increasing nutrient over-enrichment indicated by water quality deterioration, resulting in
numerous adverse effects including hypoxic zones (water with dissolved oxygen that is
too low to support marine life), species mortality, proliferation of phytoplankton and
macroalgae (including harmful algal blooms), and decreased coverage of submerged
aquatic vegetation lYBricker et al.. 2007); Appendix 101.
Indicators of nutrient enrichment may be affected by climate-related changes in estuaries
including temperature, precipitation, wind patterns, stronger estuary stratification,
increased metabolism and organic production, and sea-level rise [see discussion in
Appendix 10; (Altieri and Gedan. 2015; Statham. 2012; Rabalais et al.. 2009)1. For
example, eutrophic conditions and the extent and duration of hypoxia are predicted to
increase with changes in temperature and precipitation (Altieri and Gedan. 2015;
Rabalais et al.. 2009; Boesch et al.. 2007). Donev et al. (2012) reviewed effects of
climate change, including coastal hypoxia, on marine biodiversity and reported
population-level shifts. In estuaries affected by nutrient enrichment and
climate-associated alterations described above, conditions favor a shift in phytoplankton
composition toward greater abundance and distribution of toxic cyanobacteria associated
with an increased prevalence of harmful algal blooms (Paerl et al.. 2016a). In a
meta-analysis of temperature effects on benthic macrofauna, survival times significantly
decreased in hypoxic conditions under warmer temperatures (Vaquer-Sunver and Duarte.
June 2018
13-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
2011). Decreases in estuarine biodiversity associated with N loading can be magnified by
hydrologic factors. Glibert et al. (2014) observed a regional change in phytoplankton
community composition and decreased coverage of submerged aquatic vegetation
following a shift from long-term dry to long-term wet conditions in the early 2000s in
shallow coastal lagoons along the coast of Maryland and Virginia. Shifts in
phytoplankton community structure are known to occur in estuaries due to N enrichment,
but climate changes may at times outweigh the impacts of eutrophication (Pacrl et al..
2010).	Thus, physical changes caused by climate change should be considered when
modeling phytoplankton community response to N enrichment (Pacrl et al.. 2014).
13.3. Wetlands
Changes in mean annual temperature and frequency and magnitude of precipitation will
affect the responses of all wetlands to N loading (Appendix 11).
Increasing temperatures may strengthen N effects upon wetland ecosystems. Temperature
effects on wetlands have been demonstrated in European bog ecosystems, where
increased temperatures increased the cover of woody species and decreased Sphagnum. N
addition and temperature are known to synergistically depress Sphagnum production,
with a 1°C increase in summer temperature having an impact on the Sphagnum
equivalent to an additional 40 kg N/ha/yr (Limpens et al.. 2011).
Hydrologic regimes are important controls on wetland cycling and productivity, so
changes in the magnitude and frequency of precipitation can have strong effects on
ecosystem N retention and C storage. Increased precipitation is shown to increase
Sphagnum sensitivity to N addition-induced decreases in production (Limpens et al..
2011).	Experimental mesocosms modeling changes in precipitation to salt marshes found
that precipitation delivered in infrequent, heavy storm events decreased N retention and
plant productivity, even though storm events delivered a higher N deposition load to the
marsh (Hanson et al.. 2016; Oczkowski et al.. 2016). Shifts in precipitation towards less
frequent, more intense rain events may strengthen N deposition-induced decreases in salt
marsh N retention while weakening N deposition-induced increases in plant productivity.
June 2018
13-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
APPENDIX 14. ECOSYSTEM SERVICES
This appendix is a review of the state of the science on ecosystem services altered by N
and S deposition. The state of the science in 2008, as presented in the 2008 Integrated
Science Assessment for Oxides of Nitrogen and Sulfur-Ecological Criteria (hereafter
referred to as the 2008 ISA), is summarized and sets the foundation for the discussion on
the new literature (published from 2008-present).
14.1. Introduction
The term "ecosystem services" refers to a concept that ecosystems provide benefits to
people, directly or indirectly (Costanza et al.. 2017). and these benefits are socially and
economically valuable goods and services deserving of protection, restoration, and
enhancement (Bovd and Banzhaf. 2007). The concept of ecosystem services recognizes
that human well-being and survival are not independent of the rest of nature, humans are
an integral and interdependent part of the biosphere (Costanza et al.. 2017). In some
cases, and in line with more conventional economic thinking, ecosystem services analysis
can result in attaching monetary values to ecosystem outcomes. However, monetary
estimation may not always be possible and this does not diminish the importance of the
ecosystem services to human well-being. At a minimum, ecosystem services analysis
involves discussion and, ideally, quantification of ecological outcomes understood by
households, communities, and businesses. Explicitly linking ecosystem services with
affected people has proven difficult because of the broad definition of ecosystem services
and the numerous types of services that could be affected. An analysis of ecosystem
services specifically altered by NOx, SOx, and PM would translate the effects of ambient
concentrations and deposition into biological, physical, or monetary metrics that give
insight to public welfare effects.
The 2008 ISA documented the ecosystem services frameworks that were published at that
time and provided several examples of how N and S deposition would alter services
based on those frameworks. Since the 2008 ISA, new ecosystem services frameworks
have been developed and several new approaches have been applied to conduct
ecosystem services analysis specifically evaluating outcomes altered by N and S
deposition. The conclusions considering the full body of literature published are that
(1) there is evidence that N and S emissions/deposition have a range of effects on
ecosystem services and their social value; (2) there are some economic studies that
demonstrate such effects in broad terms; however, it remains methodologically difficult
to derive economic costs and benefits associated with specific regulatory
June 2018
14-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
decisions/standards; and (3) thousands of causal relationships are now documented
between N and S air pollution and changes in Final Ecosystem Goods and Services
(FEGS), defined as the "components of nature, directly enjoyed, consumed or used to
yield human wellbeing" (Bovd and Banzhaf. 2007). The following sections discuss the
relevant ecosystem services frameworks (Appendix 14.2). the studies in the U.S. on
ecosystem services altered by Acidification and Eutrophication (Appendix 14.3). the
studies of ecosystem services analyses conducted in countries outside of the U.S.
(Appendix 14.4 and Appendix 14.5). and a summary (Appendix 14.6).
14.2. Ecosystem Services Frameworks
Ecosystem services frameworks can help identify causal pathways between specific
ecosystem changes and potential human beneficiaries. A single classification system may
be desirable to promote standardization and communication. However, classification
systems do not by themselves solve the core issue of how to empirically develop causal
relationships between N and S and final ecosystem goods and services. Here an overview
of key frameworks is presented to give context on conceptual advances in ecosystem
services research.
Several ecosystem services frameworks were published in 2008 and described in the
2008 ISA. These frameworks provided qualitative and/or quantitative metrics by which to
identify the services that ecosystems provide to benefit human welfare and society fWRI.
2000; Costanza et al.. 1997; Daily. 1997; Pimentel etal.. 1997). It was well documented
at that time that some goods and services have explicit market value; however, the value
of other services is more difficult to assess (U.S. EPA. 2008a; Goulder and Kennedy.
1997). The most accepted framework for ecosystem services identified by the 2008 ISA
was the 2005 Millennium Ecosystem Assessment (MEA) by Hassan et al. (2005). Since
the 2008 ISA, several new ecosystem services frameworks and classification systems
have been published. These include The Economics of Ecosystems and Biodiversity
[TEEB; (Sukhdev et al.. 2014)1. the Common International Classification of Ecosystem
Services [CICES; (Haines-Young and Potschin. 2011)1. U.S. EPA Final Ecosystem
Goods and Services Classification System [FEGS-CS; (Landers and Nahlik. 2013)1. and
The National Ecosystem Services Classification System [NESCS; (U.S. EPA. 2015d)l.
The goal of the TEEB initiative is to demonstrate the economic value of services
provided by species and ecosystems and then capture the value of these services through
market- or policy-based approaches (Sukhdev et al.. 2014). The TEEB initiative uses four
primary categories of ecosystems services. Three of the four categories are from the
MEA (provisioning, regulating, and cultural services). However, TEEB replaces the
June 2018
14-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
MEA's "supporting services" with "habitat services," which emphasizes the provisioning
of species and genetic diversity.
The CICES framework was developed in part as a reaction to the increasingly diverse
ways that ecosystem services were discussed, categorized, and assessed by different
organizations around the world, the CICES program has as a primary goal to create a
more uniform approach to assessing ecosystem services (Haines-Young and Potschin.
2011). The classification approach CICES developed recognizes only final outputs from
ecosystems, in other words, "products" that are directly consumed or used by people
(Haines-Young and Potschin. 2011). Thus, the concept of "supporting services"
identified by the MEA was dropped under the assumption that these functions would be
recognized for their role in facilitating the final outputs. CICES developed a hierarchical
structure, beginning with the three broad categories (provisioning, regulating, and
cultural), which were broken down into 23 "service groups" and then 59 "service types"
(Haines-Young and Potschin. 2011).
The FEGS-CS framework uses the concept of Final Ecosystem Goods and Services
(FEGS), which are defined as a subset of ecological outcomes, specifically the
"components of nature, directly enjoyed, consumed or used to yield human wellbeing"
(Bovd and Banzhaf. 2007). The U.S. EPA FEGS-CS approach defines and classifies
338 unique FEGS, each uniquely numbered by a combination of environmental class or
subclass and a beneficiary category or subcategory (Landers and Nahlik. 2013). In
addition to FEGS-CS, the U.S. EPA also developed NESCS to specifically analyze the
human welfare impacts of policy-induced changes to ecosystems (U.S. EPA. 2015d). The
goal of NESCS is to support analysis of changes from baseline conditions, such as
cost-benefit analysis, and support systematic accounting of changes in ecosystem
services. The NESCS structure defines categories and numeric codes that are designed to
help identify flows of services from ecosystems to human beings, while distinguishing
between the producers (i.e., "supply-side") and users (i.e., "demand-side") of the service.
Lastly, Bell et al. (2017) developed the Stressor—Ecological Production function—final
ecosystem Services (STEPS) framework. STEPS produces "chains" comprised of the
biological indicator, the ecological production function (EPF, which uses ecological
components to link the biological indicator to a final ecosystem service), and the user
group who directly uses, appreciates, or values the component. To this end, STEPS is
used in conjunction with other frameworks that identify user groups, like FEGS-CS. The
framework uses a qualitative score (high, medium, low) to describe the strength of
science (SOS) for the relationship between each component in the EPF.
14-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
14.3. United States Applications
Among the ecological effects of N and S deposition, the ecosystem services altered by
acidification are the most well documented, with several new papers published since
2008 (Beier et al.. 2017; Caputo et al.. 2017; Irvine et al.. 2017; O'Dea et al.. 2017;
Banzhaf et al.. 2016; Office of Air and Radiation. 2011; Chestnut and Mills. 2005;
Banzhaf et al.. 2004; U.S. EPA. 1999; NAPAP. 1991). In the 2008 ISA, no publications
were identified describing how ecosystem services were altered as a result of N driven
eutrophication; however, several new comprehensive studies have been published since
2008 (Clark et al.. 2017; Rhodes et al.. 2017; Compton et al.. 2013; Birch et al.. 2011;
Compton et al.. 2011). Also new to the literature since 2008 is a group of publications
(Bell et al.. 2017; Clark et al.. 2017; Irvine et al.. 2017; O'Dea et al.. 2017; Rhodes et al..
2017) evaluating the FEGS altered by four modes of ecological response to N and S
deposition (aquatic acidification, terrestrial acidification, aquatic eutrophication, and
terrestrial eutrophication). In these analyses critical load exceedances for N related air
pollution were used as a model stressor from which a total of 1,104 unique chains linking
stressor to beneficiary were identified. These analyses, built using a wholistic STEPS
approach, represent a great advance for identifying the wide range of ecosystem
processes and types of final ecosystem services affected by N and S deposition. However,
it also underscores the many information gaps that exist for quantifying the dose-response
relationships between N and S air pollution exposures and final ecosystem services and
the values associated with changes in these services.
Two recently published studies (Jaramillo and Muller. 2016; Holland. 2015) have used
the Air Pollution Emissions Experiments and Policy (APEEP or AP2) to calculate the
value of damages from emissions of SO2, VOCs, NOx, PM2 5, PM10, and NH3. The model
connects changes in emissions to concentrations of several criteria air pollutants. A large
majority of the damage estimates are associated with human health impacts and
ozone-related damages to crops and forests. A very small portion of damages are linked
to forest and recreation impacts from a mix of ambient pollutants including NOx and
SOx, but this portion of the model was developed before 2006, has not been updated, and
is not well documented (i.e., the model structure is described but not the estimated
parameters).
June 2018
14-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
How do N deposition impacts oil lace lichen affect forest ecosystem services?
Native to the Pacific coast of North America, lace lichen (Ramalina memiesii [Ramalinaceae])
grows from Alaska to Baja California. It lias a unique morphology that adds to the natural
beauty of coastal woodlands along the Pacific Coast (Hastings Reserve. 2015). It is a food
source for browsing mammals, including deer, cows, and squirrels, and is also used as a nesting
material for hummingbirds and orioles (Hastings Reserve).
Photos source: 2016 AOE workshop report.
Unfortunately, lace lichen abundance is threatened by nitrogen and sulfur pollutants" deposition
(Hernandez et al.. 2016). It is very sensitive to air pollution, including exposure to nitric acid
(Riddell et al„ 2008). and it has been found to contribute to the global uptake of sulfur dioxide,
an airborne pollutant (Gries et al.. 1997). Lace lichen was formerly found in the San Jacinto
Mountains near Los Angeles and on the surrounding coastal plain but now it only grows at
elevations above the smog layer (Hastings Reserve. 2015).
Declines in lace lichen abundance can reduce ecosystem services to humans in several ways. In
addition to reducing the aesthetic benefits provided by lichens, declines in lace lichen may be
indirectly contributing to declining populations of the spotted owl, which is ail endangered,
widely recognized, and charismatic species native to Pacific Northwest forests. This indirect
impact may be occurring because the northern flying squirrel, which is an almost exclusive food
source for the spotted owl, relies exclusively on forage lichens as a winter time food source.
14.3.1. Acidification
Prior to 2008, ecosystem acidification in the U.S. was evaluated with regard to how
ecosystems services respond to lowering NOx and SOx emissions. Initial monetary
valuation of the total annual value for improvements in recreational fishing (NAPAP.
1991) were improved with updated economic studies and new models for estimating the
changes in fish stocks and catch rates in the Adirondacks U.S. EPA (1999). A few years
later, Chestnut and Mills (2005) compared the actual benefits of reducing emissions of
NOx and SOx in Title IV of the Clean Air Act Amendments (CAAA) by the estimate of
benefits made in 1990. They conclude that quantitative assessment was problematic at
that time due to a lack of measurement units to gauge changes in the quality and quantity
of ecosystem services and a lack of dose-response relationships to indicate how quality
and quantity may change as a function of changes in pollution exposures. A different
June 2018
14-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
approach, that did not rely on dose-response, documented that New York households are
willing to pay an average of $45 to $100 annually to reduce the number of acidified lakes
by 40% and improve forest health in the Adirondacks (Banzhaf et al.. 2004).
Several ecosystem services valuations for acidification published (since 2008) are given
in Table 14-1. Rea et al. (2012) paired biogeochemical modeling using the Model for
Acidification of Groundwater in Catchment (MAGIC) with two different approaches, one
benefit transfer and one random utility, to estimate ecosystem services values for
remediation scenarios of acidic deposition on lakes in New York's Adirondack Park.
Beier etal. (2017) and Caputo et al. (2017) have also calculated new monetary valuations
of acidification effects in the Adirondacks (Table 14-1). New work in the southern
Appalachian region used an approach that weds biogeochemical (e.g., MAGIC) and
economic modeling to evaluate the cost of aquatic and terrestrial acidification (Banzhaf et
al.. 2016).
Two new studies used FEGS-CS to quantify how many FEGS are altered by ecosystem
acidification and associated CL or TL exceedance (Irvine et al.. 2017; O'Dea et al.. 2017).
O'Dea et al. (2017) identified human beneficiary groups for each ecological endpoint
affected by aquatic acidification, including effects on otter, mink, loon, aquatic
vegetation, shellfish, brook trout, and bass. Irvine et al. (2017) documented the effects of
terrestrial acidification on two acid-sensitive tree species, balsam fir (Abies balsamea)
and white ash (Fraxinus americana) (Table 14-1).
Overall, the new literature since the 2008 ISA includes studies that better characterize
ecosystem services valuation by pairing biogeochemical modeling and benefit transfer
equations informed by WTP surveys, especially for the Adirondacks and Shenandoah
regions. Aside from valuation, the estimate of the total number of ecosystem services
affected by N and S deposition is more well quantified by the new studies that use
FEG-CS. However, for many regions and specific services, poorly characterized
dose-response between deposition, ecological effect, and services are the greatest
challenge in developing specific data on the economic benefits of emission reductions
(NAPAP. 2011).
14-6
DRAFT: Do Not Cite or Quote

-------
Table 14-1 Ecosystem services research related to ecosystem acidification.
Mode
Region
Effect
Reference
Aquatic
acidification
Adirondacks
In 1990, the estimates of total annual value for improvements
due to lowering NOx and SOx emissions in recreational
fishing were $12 to $24 million.
NAPAP (1991)
Aquatic
acidification
Adirondacks
Reported estimates of annual value of improvements in
conditions for recreational fishing in the Adirondacks of about
$65 million for the reductions in acid deposition as a result of
the 1990 CAAA, primarily Title IV.
U.S. EPA (1999)
Aquatic
acidification

Compare the actual benefits of reducing emissions of NOx
and SOx in Title IV of the Clean Air Act Amendments (CAAA)
by the estimate of benefits made in 1990.
Chestnut and Mills
(2005)
Aquatic
acidification
Adirondacks
New York households are willing to pay an average of $45 to
$100 annually to reduce the number of acidified lakes by 40%
and improve forest health in the Adirondacks.
Banzhaf et al.
(2004)
Aquatic	Adirondacks Linked NOx and SOx emissions estimates to CMAQ	Office of Air and
acidification	modeling, MAGIC modeling, then a Random Effects Model Radiation (2011)
that was developed to account for fishing site choices made
by recreational fishers based on attributes of sites specifically
in the Adirondack region. The difference in economic welfare
values between the value of fishable (i.e., not impaired) lakes
in the with-CAAA scenario and the without-CAAA scenario
represented the benefits to recreational fishing in the
Adirondack region associated with the CAAA.
The changes in percent base saturation levels in timber	Office of Air and
harvest areas were mapped in relation to potential changes in Radiation (2011)
the growth and health of tree species present in these areas
and the likely effects of altered tree growth and health on
timber harvest rates and volumes. Of the forest types of
interest, the paper birch forest type experiences the greatest
increase in percent base saturation due to the CAAA,
followed by the eastern hemlock and the sugar
maple/beech/yellow birch forest types.
Biogeochemical modeling (e.g., MAGIC) is combined with a Rea et al. (2012)
benefit transfer approach based on a stated preference study,
which captures ecosystem services values as a whole. Total
values range from $547 million to $1 billion per year. In
contrast, an approach based on a preference study, which
only captures recreation fishing values, estimated benefits
ranging from $7 million to $9 million per year.
Terrestrial Adirondacks
Acidification
Aquatic	Adirondacks
acidification (44 lakes)
Aquatic and Southern
terrestrial Appalachia
acidification (169 sites)
Hypothetical scenarios of lower emissions are coupled with
biogeochemical modeling (e.g., MAGIC) and a novel stated
preference survey to determine individuals' WTP is used to
generate aggregate benefits of about $3.7 billion, or about
$16 per year per household in the region.
Banzhaf et al.
(2016)
June 2018
14-7
DRAFT: Do Not Cite or Quote

-------
Table 14-1 (Continued): Ecosystem services research related to ecosystem
acidification.
Mode	Region	Effect	Reference
Terrestrial Adirondacks Blending regression analysis of tree growth and soil condition Beier et al. (2017)
acidification	with a benefit transfer approach, it was estimated that
acid-impaired hardwood forests provide roughly half of the
potential benefits of forests on moderate to well-buffered
soils—an estimated loss of-$10,000 ha-1 in net present
value of wood products, maple syrup, carbon sequestration,
and visual quality.
Aquatic	Adirondacks Acidic deposition has had little effect on lakes water suitable Beier et al. (2017)
acidification (52 study for drinking in the region. However, as pH declines in lakes,
lakes)	the estimated value of recreational fishing decreases
significantly due to loss of desirable fish. The expected value
of recreational fishing increased with increasing pH, from a
minimum of $4.41 angler/day in unstocked, acid-impaired
lakes to a maximum of $38.40 angler/day in well-buffered
lakes that were stocked with trout.
Aquatic	Adirondacks Analysis of benefit transfer using willingness-to-pay (WTP) for Caputo et al. (2017)
acidification	recreational fishing and logistic regression offish populations
demonstrated that under low emissions scenarios combined
with fish stocking. These are the same results reported in
Beier et al. (2017).
Aquatic	U.S. national- Using FEGS-CS to quantify the total number of causal	O'Dea et al. (2017)
acidification scale	relationships between aquatic acidification driven by N and S
deposition and beneficiaries. Effects were documented on
otter, mink, loon, aquatic vegetation, shellfish, brook trout,
and bass.
Terrestrial Northeastern Using FEGS-CS to identify 160 chains with 10 classes of Irvine et al. (2017)
acidification U.S.	human beneficiaries for balsam fir and white ash combined,
the authors concluded there are resources at risk that the
public may value. Estimated ~1,200 chains if all vulnerable
tree species are evaluated.
14.3.2. Eutrophication
1	Since the 2008 ISA, several comprehensive studies have been published on the
2	ecosystems services related to N pollution in the U.S. (Table 14-4). These include an
3	evaluation of services effected by multiple N inputs (including N deposition) to the
4	Chesapeake Bay (Birch etal.. 2011) and syntheses by Compton et al. (2011) and
5	Compton et al. (2013) of the cost-benefits on N loading across the nation for services that
6	included coastal fish harvests, recreational uses of inland and coastal waters, lakefront
7	property values, and water treatment costs. Further expanding on this work, Sobotaetal.
8	(2015) calculated the amount of N loading from human activity that is leaked into
9	adjacent ecosystems and the associated effects on ecosystem services. In their work,
10	Sobotaetal. (2015) specifically identified the costs of the atmospheric portion of total N
June 2018
14-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
loading. (Baron et al.. 2012) evaluated the interactions between climate and N,
identifying some key services in which a changing climate will modify the response to N.
Birch etal. (2011) examined the spatial and temporal movement of reactive N through
multiple ecosystems and media in the Chesapeake Bay watershed and estimated damage
costs in several categories (Figure 14-1). Their analysis estimates the contribution of
atmospheric N sources to the total annual N loads delivered to the bay estuary. In other
words, roughly 20% of N related ecosystem services impacts in the estuary are
attributable to atmospheric N deposition in the watershed.
Building on the work by Birch etal. (2011). a review and synthesis of the published
literature by Compton et al. (2011) compared estimates of the average damage costs/unit
of N across a range of affected services. Furthing building on this work, Sobotaetal.
(2015) estimated how N leaked to the environment (air/deposition, surface freshwater,
groundwater, and coastal zones) alters services by multiplying watershed-level N inputs
(8-digit U.S. Geologic Survey hydrologic unit codes; [HUC8s]) with published
coefficients describing nutrient uptake efficiency, leaching losses, and gaseous emissions.
Sobotaetal. (2015) then applied a per-unit damage cost values ($/N) for a wide variety
of ecological effects (based on estimates from various sources, Table 14-2) to the leakage
estimates.
14-9
DRAFT: Do Not Cite or Quote

-------
Why are aquatic grasses on the path between N deposition and ecosystem services in the
Chesapeake Bay?
The Chesapeake Bay is the largest estuary (4,480 sq miles) in the U.S. draining from eight major river
basins covering 64,000 sq miles of watershed in six states and a population of 18 million watershed
residents. The public's use of the bay is significant with more than 700 public access points on the bay and
its tributaries. More than 3,600 species of plants and animals are supported by the bay, including over
80 resident and migratory waterfowl. About 500 million pounds of seafood is produced annually by the
bay (https://www.chesapeakebay.net).
Unfortunately, this vital natural resource and estuarine ecosystem are under stress, and nitrogen deposition
in the Chesapeake Bay watershed is one of the main contributing sources. The Chesapeake Bay Program
reports that one-third of the nitrogen loading the bay comes from atmospheric deposition, often as nitrogen
oxides or ammonia (https://www.chesapeakebav.net/issues/air pollution). Nitrogen air emission sources
include automobiles, fossil fuel-fired electric generating units, and intensive animal operations both within
and upwind of the bay watershed.
One of the most important indicators of environmental stress to the Chesapeake Bay is the loss of
submerged aquatic vegetation (SAV). Nutrient over-enrichment and sediment overloading the bay has led
to SAV loss. Although SAV appears to be recovering as NOx emissions have declined in recent decades,
the bay still contains only an estimated 53% of the SAV it once supported. The Chesapeake Bay
Program's target restoration goal is 185,000 acres
fhttps://www.chesapeakebay.net/state/uriderwater grasses).
Photo courtesy of Chesapeake Bay Program
SAV is a critical indicator of the bay's health in part because of the essential role it plays in supporting a
wide range of final ecosystem services to humans. The 80,000 acres of underwater grasses protect and
host young and molting blue crabs and juvenile, for example
(https://wwyv.chesapeakebav.net/discover/facts). SAV supports natural ecosystems in additional ways,
including serving as a food source for wildlife, waterfowl, and humans and as habitat (and protection) for
wildlife and waterfowl (Barbier et al.. 2011; Kemp et al.. 2005) (Johnston et al- 2002). In turn, SAV
facilitates important recreation, tourism, and commerce, including wildlife and waterfowl watching,
fishing, and crabbing. In addition to these valuable indicators of overall WQ and ecosystem health, SAV
deters and dissipates wave energy and promotes sedimentation for landowners along the bay, thus,
mitigating erosion and protecting and sustaining land value.
The SAV connection is, therefore, one important pathway through which N deposition affects the well-
being of human populations in and around the bay.
June 2018
14-10
DRAFT: Do Not Cite or Quote

-------
Chemical Nitrogen Cascade: Chesapeake Bay Watershed
I! mi Villi in

Utilmes
*2,0M
In&uniy
«,WM1
Mobile S.UL6
170,0X1
<3iie7 iixircev
14,DM
r
SH,_ Umbiinm

Mm-Afjiwltirf
KjOflQ
FIGURE 1. Chemical Nriroqen Cascade in tfce Chesapeake Bny Walenkts" IteucvyeBrt- S« S3 for s-oinces and colculaticMK.
N = nitrogen; N.E. = no estimate available; N20 = nitrous oxide; NH3 = ammonia; NOx = nitrogen oxide.
Source: Birch et al. (2011).
Figure 14-1 Economic nitrogen cascade in the Chesapeake Bay Watershed.
June 2018
14-11
DRAFT: Do Not Cite or Quote

-------
Table 14-2 Potential damage costs of nitrogen (N) ($/kg N; 2008 or as reported)
to air, land, and water resources in the conterminous U.S. in the early
2000s as synthesized by Sobota et al. (2015). Low, median, and high
costs derive from the specific damage cost reference. Negative
values indicate an economic benefit.
Cost ($/kg N)
N damage type
System
Low
Median
High
Reference
From atmospheric NOx
Increased incidence of
respiratory disease
Air/Climate
12.88
23.10
38.63
Van Grinsven et al. (2013)
Birch et al. (2011)
Declining visibility—loss of
aesthetics
Air/Climate
0.31
0.31
0.31
Birch et al. (2011)
Increased effects of airborne
particulates/increased carbon
sequestration in forests
(includes benefits)
Air/Climate
-11.59
-4.51
2.58
Van Grinsven et al. (2013)
Increased damages to
buildings from acid
Land
0.09
0.09
0.09
Birch et al. (2011)
Increased ozone exposure to
crops
Land
1.29
1.51
2.58
Van Grinsven et al. (2013) Van
Grinsven etal. (2013): Birch et
al. (2011)
Increased ozone exposure to
forests
Land
0.89
0.89
0.89
Birch et al. (2011)
Increased loss of plant
biodiversity from N
enrichment
Land
2.58
7.73
12.88
Van Grinsven etal. (2013)
From atmosphere NH3
Increased incidence of
respiratory disease
Air/Climate
2.58
4.93
25.75
Van Grinsven et al. (2013) Van
Grinsven etal. (2013): Birch et
al. (2011)
Declining visibility—loss of
aesthetics
Air/Climate
0.31
0.31
0.31
Birch etal. (2011)
Increased effects of airborne
particulates/increased carbon
sequestration in forests
(includes benefits)
Air/Climate
-3.86
-1.93
-1.93
Van Grinsven et al. (2013)
Increased damages to
buildings from particulates
Land
0.09
0.09
0.09
Birch et al. (2011)
June 2018
14-12
DRAFT: Do Not Cite or Quote

-------
Table 14-2 (Continued): Potential damage costs of nitrogen (N) ($/kg N; 2008 or as
reported) to air, land, and water resources in the
conterminous U.S. in the early 2000s as synthesized by
{Sobota, 2015, 2944570@@author-year}. Low, median,
and high costs derive from the specific damage cost
reference. Negative values indicate an economic benefit.



Cost ($/kg N)


N damage type
System
Low
Median
High
Reference
Increased loss of plant
biodiversity
Land
2.58
7.73
12.88
Van Grinsven etal. (2013)
From N2O
Increased ultra-violet light
exposure from
ozone—humans
Air/Climate
1.29
1.33
3.86
Van Grinsven et al. (2013)
Compton et al. (2011)
Increased emission of
greenhouse gas
Air/Climate
5.15
13.52
21.89
Van Grinsven et al. (2013)
Increased ultra-violet
exposure from ozone—crops
Air/Climate
1.33
1.33
1.33
Birch etal. (2011)
From surface freshwater N loading
Declining waterfront property
value
Freshwater
0.21
0.21
0.21
Dodds et al. (2009)
Loss of recreational use
Freshwater
0.17
0.17
0.17
Dodds et al. (2009)
Loss of endangered species
Freshwater
0.01
0.01
0.01
Dodds et al. (2009)
Increased eutrophication
Freshwater
6.44
16.10
25.75
Van Grinsven etal. (2013)
Compton et al. (2011)
Undesirable odor and taste
Drinking
water
0.14
0.14
0.14
Kusiima and Powers (2010)
Nitrate contamination
Drinking
water
0.54
0.54
0.54
Compton et al. (2011)
Increased colon cancer risk
Drinking
water
1.76
1.76
5.15
Van Grinsven etal. (2013)
From groundwater N loading
Undesirable odor and taste
Drinking
water
0.14
0.14
0.14
Kusiima and Powers (2010)
Nitrate contamination
Drinking
water
0.54
0.54
0.54
Compton et al. (2011)
Increased colon cancer risk
Drinking
water
1.76
1.76
5.15
Van Grinsven etal. (2013)
June 2018
14-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Table 14-2 (Continued): Potential damage costs of nitrogen (N) ($/kg N; 2008 or as
reported) to air, land, and water resources in the
conterminous U.S. in the early 2000s as synthesized by
{Sobota, 2015, 2944570@@author-year}. Low, median,
and high costs derive from the specific damage cost
reference. Negative values indicate an economic benefit.



Cost ($/kg N)


N damage type
System
Low
Median
High
Reference
From coastal N loading
Loss of recreational use
Coastal
zone
6.38
6.38
6.38
Birch et al. (2011)
Declines in fisheries and
estuarine/marine habitat
Coastal
zone
6.00
15.84a
26.00
Van Grinsven et al. (2013) Van
Grinsven et al. (2013):
ComDton et al. (2011)
aExcludes $56/kg N from submerged aquatic vegetation loss in the Gulf of Mexico from Compton et al. (2011).
Still another advance is offered by Clark et al. (2017)and Rhodes et al. (2017). who
synthesized information on N induced terrestrial and aquatic eutrophication, respectively,
from the published literature to link the ecological effects of critical load exceedances
with human beneficiaries by using STEPS and FEGS-CS. Rhodes et al. (2017) identified
that N loading to aquatic ecosystems affects 154 chains that link changes in biological
indicators of aquatic eutrophication (a shift in phytoplankton community) to FEGS,
13 ecological production functions (EPF) within three different ecosystems (alpine lakes,
lakes, and estuaries) and 18 classes of human beneficiaries that potentially will be
effected by a change in one of these endpoints. For terrestrial eutrophication, Clark et al.
(2017) identified that N critical load exceedences affected beneficiary types through
582 individual chains in the five ecoregions examined (Eastern Temperate Forests,
Marine West Coast Forests, Northwestern Forested Mountains, North American Deserts,
Mediterranean California) and 66 FEGS across a range of final ecosystem services
categories (21 categories; e.g., changes in timber production, fire regimes, and native
plant and animal communities) (Figure 14-3 and Table 14-3).
June 2018
14-14
DRAFT: Do Not Cite or Quote

-------
Level I CEC ecoregions
Eastern Temperate Forests	North American Deserts
Marine West Coast Foreste M Northwestern Foisted Mountains
Mediterranean California	01 tier Ecoregions
Map of (a) Level 1 ecoregions examined and (b) the count of unique chains, biological indicators, Final Ecosystem Goods and
Services, and beneficiaries for each ecoregion.
Source: Clark et al. (2017).
Figure 14-2 Map of ecosystem services altered by nitrogen critical load
exceedance.
June 2018
14-15
DRAFT: Do Not Cite or Quote

-------
Table 14-3 Numbers of chains, Final Ecosystem Goods and Services (FEGS),
and beneficiaries (bens) associated with each initial biological
indicator Clark et al. (2017).
Biological Indicator
No. Chains
No. FEGS
No.
Benefits
Increased grass-to-forb ratio and/or increase in total biomass
192
11
12
Decreased lichen biodiversity
62
7
11
Decreased native mycorrhizal diversity
61
5
10
Decreased abundance of ectomycorrhizal fungi
43
3
12
Increased bark beetle abundance
34
4
12
Decreased survival of bigtooth aspen
32
4
11
Decreased survival of scarlet oak
32
4
11
Decreased survival of trembling aspen
32
4
11
Decreased abundance of arbuscular mycorrhizal fungi
25
3
10
Decreased growth of red pine
25
3
10
Increased cover of understory nitrophilic species
18
2
10
Increased bacteria-to-fungi ratio
17
2
9
Change in herbaceous community composition
7
1
7
Increase in N leaching
2
1
2
FEGS = Final Ecosystem Goods and Services.
June 2018
14-16
DRAFT: Do Not Cite or Quote

-------
Table 14-4 Ecosystem services research related to Nitrogen (N)-driven
eutrophication.
Reference
Mode	Region	Effect	HERO ID
Estuarine N Chesapeake Recreation and commercial fishing are the main ecosystem Birch et al. (2011)
driven	Bay	services impacts; however, it is difficult to quantify these
eutrophication	impacts due to data limitations. The analysis finds that,
although atmospheric releases of N are lower than direct
releases to land and water, their total damages are larger yet
have associated abatement costs that are relatively low.
Includes	Damage to services associated with productivity, biodiversity, Compton et al.
acidification	recreation, and clean water are less certain and although (2011)
and	generally lower, these costs are quite variable (<$2.2-$56
eutrophication	per kg N). In the current Chesapeake Bay restoration effort,
for example, the collection of available damage costs clearly
exceeds the projected abatement costs to reduce N loads to
the Bay.
Includes	National	Annual damage costs associated with anthropogenic N
acidification scale,	leakage range from $1.94 to $2,255/hectare per year,
and	calculated for Nationally, the total quantifiable damages were estimated to
eutrophication 8-digit HUCs be between $81-$144 billion per year. Between 14-24% of
the potential damage costs were associated with fossil fuel
combustion. Areas with the largest damage costs
corresponded to areas with the largest N inputs and
leakages, such as the upper Midwest and central California.
Sobota et al. (2015)
Terrestrial N National scale N critical load exceedences affected beneficiary types	Clark et al. (2017)
driven	through 582 individual and 66 FEGS across a range of final
eutrophication	ecosystem services categories.
Aquatic N National scale Using STEPS and FEGS-CS, the authors identify 154 chains Rhodes et al. (2017)
driven	that link changes in biological indicators of aquatic
eutrophication	eutrophication (a shift in phytoplankton community) to FEGS,
and 18 classes of human beneficiaries.
14.3.3. Nitrogen and Climate Modification
1	Some work has discussed the implications of climate change and N loading on ecosystem
2	services such as fish harvests, property values, water treatment, human and wildlife
3	health, ocean acidification, and freshwater diversity (Baron et al.. 2012). The combined
4	effect of N loading and climate change on the economic value of water resources and
5	related products has yet to be evaluated, and even the separate economic effects of N
6	loading or climate change are difficult to determine. However, climate change will
7	modify the effects of N loading on some key ecological services. Porter etal. (2013) and
June 2018	14-17	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
Compton et al. (2013) also identified some effects on the interactions between nitrogen
and climate, which include:
•	As eutrophication increases with warmer water temperatures, there will be costs
associated with upgrades of municipal drinking water treatment facilities, the
purchase of bottled water, and the health costs of NO, in drinking water leading
to toxicity and disease (Compton et al.. 2011).
•	The U.S. will increasingly rely on groundwater for drinking water under future
climate change scenarios (U.S. Global Change Research Program. 2009). creating
a strong potential for increased costs for treating exposure to NOs -stimulated
disease.
•	Increasingly, N enrichment is correlated in waters with pathogen abundance and
human and wildlife diseases (Johnson et al.. 2010). Climate warming opens the
possibility for more vector-transmitted diseases to migrate to higher latitudes,
where N loading may enhance their success (Johnson et al.. 2010).
•	Acidification causes direct harm to calcifying shellfish and crustaceans (Howarth
et al.. 2012). Changes in climate and the N cycle will intensify ocean acidification,
and there are feedbacks from acidification to N cycling (Donev et al.. 2009).
Table 14-5 provides a summary of recent studies and findings regarding the
economic effects of ocean acidification on U.S. fisheries.
14-18
DRAFT: Do Not Cite or Quote

-------
Table 14-5 Summary of recent literature examining economic impacts of ocean
acidification on U.S. fisheries.
Region
Type of Fish
Key Findings
Reference
Northeast
Crustaceans
Ocean acidification alone results in very minor changes in
Ainsworth et al.
Pacific from
(especially
landings relative to the 2060 baseline projections; however,
(2011)
northern
shrimp),
ocean acidification in combination with three other climate

California to
echinoderms,
change effects (primary productivity, zooplankton

southeast
molluscs, and
community structure, and ocean deoxygenation) reduces

Alaska
euphaeids
landings by about 20%, which suggests that synergies exist.

U.S.
Sea scallops
Increased scallop growth rates from warming predicted to
Coolev et al.


outweigh decreasing growth rates from ocean acidification
(2015)


until 2030, at which point the negative influence of ocean



acidification will become the dominant effect.



After 2030, fewer scallops will attain largest ("U10")size



before they are harvested in current fishing levels are



maintained.

U.S.
Various
Considerable revenue declines, job losses, and indirect
Coolev and

commercial
economic costs are possible due to ocean acidification
Donev (2009)

species with a
broadly damaging marine habitats, altering marine resource


focus on molluscs
availability, and disrupting other ecosystem services.



Under a moderate 2% discount rate, U.S. ex-vessel revenue



losses are predicted to be $0.6-$2.6 billion and broader



economic losses are predicted to be $1.5-$6.4 billion.

U.S.
Shellfish
Analysis of sensitivity and adaptive capacity to OA across
Ekstrom et al.


23 regions indicates that the most socially vulnerable
(2015)


communities can be found on the U.S. East Coast and Gulf



of Mexico.



The East Coast is vulnerable mostly due to economic



dependence whereas the Gulf of Mexico is vulnerable



because of low adaptive capacity.

Washington,
Shellfish
Survey results from a sample of shellfish industry
Mabardv et al.
Oregon,

participants (70% owners or managers of hatcheries)
(2015)
California

indicated that about half had experienced negative impacts



from OA and that more than half the industry felt they would



be somewhat or definitely able to adapt to OA.

Global,
Shellfish
The economic cost of mollusc loss estimates for the U.S.
Narita et al.
disaggregated

are estimated to be roughly $400 million USD, assuming no
(2012)
by region

change in income but taking into account welfare losses



from shellfish price changes.

14.4. European and Canadian Applications
Since the 2008 ISA, several qualitative and quantitative assessments of either N
deposition or the combination ofN and S deposition have been conducted in the U.K.
June 2018
14-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
(Jones et al.. 2014). across Europe (dc Vries et al.. 2014b; de Vries et al.. 2014c; Van
Grinsven et al.. 2013). and Canada (Ahcrnc and Posch. 2013). Although these
assessments have varied considerably in their approaches, all have used simplified
methods that intentionally omitted much of the mechanistic and spatial complexity in
how deposition affects ecosystem services. Some of the qualitative and conceptual
assessments provided descriptions of impacts of N deposition on ecosystems (de Vries et
al.. 2014b; de Vries et al.. 2014c; Jones et al.. 2014).
In Canada, Aherne and Posch (2013) examined critical load exceedances to assess the
impacts of N and S deposition on water quality (eutrophication), soil quality
(acidification), and plant species diversity. The approach was quantitative, but did not
assign values for ecosystem services. Instead, the analysis focused on the sources of these
emissions and understanding what regions within Canada faced a loss of these services
because critical loads had been exceeded.
Studies in Europe found that N deposition has both positive and negative effects on
ecosystem services, de Vries et al. (2014b) modeled ecosystem response to N deposition,
using the Very Simple Dynamic soil model. The model quantified changes in these
services for the 2000-2050 period using three emissions scenarios: 1980 emissions,
current legislation, and a scenario where deposition does not exceed critical loads for any
ecosystem. The biggest differences among the emission scenarios were that regulations
imposed since the 1980s created significant decreases in areas receiving critical N loads
for water NOs and Al, but also likely decreased CO2 uptake by -20%. A second study by
Van Grinsven et al. (2013) conducted economic analysis of the influence of N cycling in
Europe and found that most economic impacts were related to atmospheric N emissions.
Although there were positive impacts on agricultural production from fertilization, there
were large negative impacts of atmospheric N emissions on human health (direct effects
on human health are outside the scope of the secondary NAAQS review) and ecosystem
function. Overall, the costs of added reactive N (atmospheric and agricultural) in the
European Union were between Ł75 and Ł485 billion/year. Within the U.K., Jones et al.
(2014) conducted a quantitative analysis that focused on six main ecosystem services
impact pathways (Figure 14-3): timber production, livestock production, carbon
sequestration, greenhouse gas (GHG) emissions, recreational fishing, and biodiversity
appreciation. For the first three pathways, declining N deposition is estimated to result in
ecosystem services losses of Ł27.2 million/year, due to decreased fertilization of
woodlands, grasslands, and heathland. For the other three pathways, declining deposition
is estimated to result in ecosystem services gains totaling Ł93 million/year, due to
decreased GHG emissions, improved water quality for recreational fishing, and enhanced
nonuse values for biodiversity through habitat protection. The estimated net gains to U.K.
society are, therefore, Ł65.8 million/year, with a 95% confidence interval of
14-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
Ł15.1—Ł123.2 million pounds/year. Because of the large number of assumptions and
simplifications, these studies present a broad conceptual analyses demonstrating the
potential of ecosystem services to quantify the impacts of reactive N in ways that could
be valuable for policymakers, rather than as definitive assessments.
Benefltsfcosts from declining nitrogen deposition, 1987-2007
200
-100
Lower bound
estimate
Central estimate
Upper bound
estimate
C02 = carbon dioxide; N20 = nitrous oxide; recr = recreational.
Source: Jones et al. (2014).
Biodiversity
¦	Recr. fishing
N20 emissions
¦	C02 sequestration
¦	Livestock
¦	Timber
Figure 14-3 Benefits and costs associated with the 25% decline in nitrogen
deposition in the U.K. since 1990.
14.5. Global Perspective
There is one study on ecosystem services with a global perspective published since 2008
and one study published in 2005 not included in the 2008 ISA. Alcamo et al. (2005)
modeled global and regional impacts of stressors that included critical load exceedance
for acidification and eutrophication on provisions, regulations, support, and cultural
ecosystem services. However, these impacts represent the combined effects of eight
drivers (including climate change, land use change, mineral extractions, and N fertilizer
use). Consequently, the analysis does not define ecosystem services changes that are
specifically attributable to changes inNorS deposition.
June 2018
14-21
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Erisman et al. (2013) examined the impact of humans on the global N cycle and how
these impacts relate to changes in ecosystem services. They group services into the
following categories:
•	Drinking water (nitrates),
•	Air pollution effects on human health,
•	Air pollution effects on crops (due to ozone),
•	Freshwater eutrophication (defined as areas where the concentration of nitrate
exceeds 1 mg NO3 -N/L) (UNEP. 2007; Camargo and Alonso. 2006; Vorosmartv
et al.. 2005).
•	Biodiversity loss (they identify the global critical load for biodiversity loss as
5-10 kg/ha/yr) (Bobbink et al.. 2010),
•	Stratospheric ozone depletion,
•	Changes in climate, and
•	Coastal dead zones.
It is especially noteworthy that the authors selected thresholds of level of N addition to an
ecosystem which begins to cause adverse effects. The paper concludes that better
quantitative relationships need to be established between N and the effects on ecosystems
at smaller scales, including a better understanding of how N shortages can affect certain
populations.
14.6. Summary
Since 2008, several studies have identified a range of ways in which N-S deposition
could affect socially valuable ecosystem services. Several new ecosystem services
frameworks and classification systems have been published, including The Economics of
Ecosystems and Biodiversity [TEEB; (Sukhdev et al.. 2014)1. the Common International
Classification of Ecosystem Services [CICES; (Haines-Young and Potschin. 2011)1. and
U.S. EPA Final Ecosystem Goods and Services Classification System [FEGS-CS;
(Landers and Nahlik. 2013)1.
For acidification, the literature since the 2008 ISA includes studies that better
characterize ecosystem services valuation by pairing biogeochemical modeling and
benefit transfer equations informed by WTP surveys, especially for the Adirondacks and
Shenandoah regions (Table 14-1). Aside from valuation, the estimate of the total number
of ecosystem services affected by N and S deposition is more well quantified by the new
studies that use FEG-CS. However, for many regions and specific services, poorly
characterized dose-response between deposition, ecological effect, and services are the
June 2018
14-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
greatest challenge in developing specific data on the economic benefits of emission
reductions (NAPAP. 2011).
In the 2008 ISA there were no publications specifically evaluating the effects of N
deposition on ecosystem services associated with N driven eutrophication. Since the 2008
ISA, several comprehensive studies have been published on the ecosystems services
related to N pollution in the U.S. (Table 14-4). These include an evaluation of services
effected by multiple N inputs (including N deposition) to the Chesapeake, a syntheses of
the cost-benefits on N loading across the nation, and analysis of the amount of N leaked
from its intended use and the effects on ecosystem services. In their work (this work
specifically identified the costs of the atmospheric portion of total N loading). Aside from
valuation, the estimate of the total number of ecosystem services affected by N is more
well quantified by the new studies that use FEG-CS (Bell et al.. 2017; Clark et al.. 2017;
Irvine et al.. 2017; O'Dea et al.. 2017; Rhodes et al.. 2017). In these analyses critical load
exceedances for N related air pollution were used as a model stressor from which a total
of 1,104 unique chains linking stressor to beneficiary were identified.
In the time since the 2008 ISA, several qualitative and quantitative assessments of either
N deposition or the combination of N and S deposition have been conducted in the U.K.,
across Europe, Canada, and at the global scale. Although these assessments have varied
considerably in their approaches, all were intentionally simplified to neglect much of the
mechanistic and spatial complexity on how N deposition affects ecosystem services. At
the global scale, an analysis selected general quantitative thresholds ofN loading for the
onset of ecologically adverse effects. However, the paper concludes that better
quantitative relationships need to be established between N and the effects on ecosystems
at smaller scales.
The conclusions considering the full body of literature are that (1) there is evidence that
N and S emissions/deposition have a range of effects on U.S. ecosystem services and
their social value; (2) there are some economic studies that demonstrate such effects in
broad terms; however, it remains methodologically difficult to derive economic costs and
benefits associated with specific regulatory decisions/standards; and (3) thousands of
causal relationships are now documented between N and S air pollution and changes in
Final Ecosystem Goods and Services. There are resources in the literature that would be
useful in quatification and valuation, but would require additional analyses to connect to
N and S deposition.
14-23
DRAFT: Do Not Cite or Quote

-------
14.7. Supplemental Materials: Ecosystem Services Profiles of
Select Species
1	Several species profiles have been created to better characterize the ecosystem services
2	provided by species affected by NOy and SOx. The species included here are identified as
3	threatened and endangered with N identified as a contributing stressor (Hernandez et al..
4	2016). The species are: balsam fir (Abies balsamea), eelgrass (Zostera marina), green
5	turtle (Chelonia mydas), white ash (Fraxinus americana), and lace lichen (Ramalina
6	menziesii). These profiles identify the geographic distribution within the U.S., ecological
7	function, Class I areas, FEGS, and cultural importance; information on economic
8	valuation is presented when available.
14.7.1. Balsam Fir
9	Scientific Name: Abies balsamea.
10	• Family: Pinaceae (pine).
11	Symbolic Role: NA.
12	Federal or State Threatened or Endangered Species Listing Status: Endangered
13	species in Connecticut.
14	Geographic Range/Distribution:
15	• Native and Current Range: The range of balsam fir is North America, including
16	most of eastern Canada. In the U.S., it extends from Minnesota to Maine and
17	south into central Pennsylvania and the mountains of Virginia and West Virginia.
18	Of the two main varieties—balsamea and phanerolepis—it is the latter that is
19	found in Virginia and West Virginia.
20	Overlap with Class I Areas:
21	• http://www.epa.gov/visibilitv/maps.html.
22	• http://www.cpa.gov/ttn/oarpg/t 1 /fir notices/classimp. gif.
23	Habitat: Grows best in areas characterized by cool temperatures and moist soils. In the
24	U.S., it is most commonly found in mixed stands, where associates include red spruce,
25	black spruce, paper birch, aspen, and red maple (Uchvtil. 1991; USDA. 1990b).
26	Primary Threats:
27	• Related to N or S Deposition: none
28	• Other: Balsam fir are particularly vulnerable to insect damage from spruce
29	budworms (Uchvtil. 1991) and from decay caused by the red heart fungus
30	[Haematostereum sanguinolentum (USDA. 1990b)I.
June 2018	14-24	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Ecosystem Role and Function:
•	Nutrition: Provides some nutrition for mice, voles, red squirrels, and grouse
OJSDA. 1990b).
•	Cover: Provides winter cover to ungulates, including white-tailed deer and
moose, and to grouse and songbirds (Uchvtil. 1991). Also provides cover for
small mammals including martens and snowshoe hares (Darvcau et al.. 2001; de
Bellefeuille et al.. 2001; Darveau et al.. 1998).
•	Other: As with other tree species in general, balsam fir trees provide a number of
other intermediate ecosystem services that are indirectly valued by humans
including nutrient cycling, carbon sequestration, soil stabilization and erosion
control, water regulation and filtration, and air pollutant filtration (Kricgcr. 2001).
Final Ecosystem Services:
•	Direct Human Uses:
•	Raw material in wood products and other industries: Used as pulpwood
and lumber for a variety of products. Oleoresin from balsam fir bark blisters
has various uses including as a medium for mounting microscopic specimens
and cement for optical systems (USDA. 1990b). It is also used in developing
fragrances (Zcrbc et al.. 2012).
•	Energy/fuel source: Wood wastes for producers using balsam fir wood or
pulp are sometimes used for energy (USDA. 1990b).
•	Cultural and human health. Balsam fir are commonly grown and used as
Christmas trees and as material for wreath-making. In 2009 they accounted
for 7% of Christmas tree sales in the U.S. (USDA. 2010). A large number of
ethnobotanical uses of balsam fir by Native Americans has also been
documented. The Native American Ethnobotany database maintained by
University of Michigan (2015) references documents citing 87 specific uses
by 13 different tribes, including a wide variety of medicinal uses
(e.g., dermatological and venereal aid; cold and cough remedy) and uses as a
material for rugs and bedding, sewing, and canoe waterproofing.
14.7.2. Eel Grass
Scientific Name: Zostera marina (Zosteraceae).
Symbolic, Federal or State Threatened or Endangered Species Listing Status: Listed
as a conservation species of least concern.
Geographic Range/Distribution:
• Native Range: Eelgrass is native to coastal regions of both the Pacific and
Atlantic Oceans along the coast of North America.
June 2018
14-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
•	Current Range: Eelgrass is currently found in coastal regions in the Pacific
Ocean along the western U.S. and Canada. It is also found along the eastern coast
of North America from Canada to the Mid-Atlantic region in the U.S. (LJSDA.
2015a). Seagrasses have been disappearing from their native range at an average
rate of 110 km2/year since 1980 (Wavcott et al.. 2009) and are especially sensitive
to changes in temperature along the southern limits of their current range in North
America (Kaldv. 2014).
Overlap with Class I Areas:
•	http://www.epa.gov/visibilitv/maps.html.
•	http://www.cpa.gov/ttn/oarpg/t 1 /fir notices/classimp. gif.
Habitat: Coastlines in both the Atlantic and Pacific oceans. It can be found in bays,
estuaries, lagoons, and beaches.
Primary Threats:
•	Related to N or S Deposition: (Hernandez et al.. 2016).
•	Other: Threats to eelgrass include dredging for coastal development, overfishing
of predator species that lead to losses of herbivores that clean the leaves, and
sediment and nutrient loading from coastal development (Wavcott et al.. 2009;
Orth et al.. 2006). More broad changes in climate, such as temperature and
sea-level rise also contribute to the loss of seagrasses (Kaldv. 2014; Douglass et
al.. 2010).
Ecosystem Role and Function: Eelgrass are an important marine species for primary
productivity (Duartc. 2013). and provides a suitable reproductive habitat and nursery
grounds for many species of fish and shellfish (Barbier et al.. 2011). Birds, invertebrates,
green turtles, and manatees use eelgrass as a food source. Some commercially viable fish
species like pacific salmon, pacific herring, and Dungeness crabs rely on eelgrass to
provide cover and habitat (Plummer et al.. 2013; Orth et al.. 2006). Another function of
eelgrass is its ability to improve water quality through nutrient uptake and particle
deposition. Eelgrass is an important species for coastal erosion protection because it helps
control wave attenuation (Barbier et al.. 2011). Eelgrass also provides indirect human
benefits through carbon sequestration because it uses dissolved carbon in seawater to
grow, and at the end of its lifecycle, the carbon is buried and stored as detritus (Barbier et
al.. 2011).
Final Ecosystem Services:
•	Direct Human Uses:
•	Cultural: Native Americans used dried eelgrass to bake into cakes and also
for smoking deer meat (Felger and Moser. 1985). Native Americans consume
the roots of the plant or use it to flavor other foods (University of Michigan.
2015).
•	Raw Material: Eelgrass is used to pack crabs to keep them moist during
transport in the Mid-Atlantic region of the U.S. (Barbier et al.. 2011).
14-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
14.7.3. Green Turtle
Scientific Name: Chelonia mydas (Cheloniidae).
Symbolic Role: Green sea turtles are symbols of guidance and protection in Hawaiian
culture.
Federal or State Threatened or Endangered Species Listing Status: Listed as an
endangered species (Scminoff. 2004).
Geographic Range/Distribution:
•	Native Range: Green sea turtles have distinct native populations in both the
Pacific and Atlantic oceans.
•	Current Range: In the North American Atlantic, nesting sites are found in
coastal regions in Florida, as well as Georgia, South Carolina, and North Carolina
(IJSFWS. 2015). In the Pacific, green turtles nest along the French Frigate Shoals
in the northwestern Hawaiian Islands (NQAA Fisheries Pacific Islands Regional
Office. 2015a).
Overlap with Class I Areas:
•	http://www.epa.gov/visibilitv/maps.html.
•	http://www.epa.gov/ttn/oarpg/t 1 /fir notices/classimp. gif.
Habitat: Mature green sea turtles are found in shallow coastal regions with large
seagrass beds such as bays, reefs, and inlets. Green turtles nest on open beaches with
minimal disturbance (IJSFWS. 2015).
Primary Threats:
•	Related to N or S Deposition: (Hernandez et al.. 2016).
•	Other: Other major threats to green sea turtles include entanglement in fishing
gear, incidental by catch, and collisions with fishing boats or jet skis. Green turtles
also face threats from illegal human harvesting and habitat loss and degradation
due to human development along coastal areas and reefs (Malama na Honu.
2015).
Ecosystem Role and Function: Green sea turtles are herbivores that mostly feed on
seagrasses and algae in near-shore marine ecosystems. Green sea turtle eggs are a food
source for many coastal species like ghost crabs or marine birds, as well as other
scavenging animals. Tiger sharks are the only nonhuman predator for large juvenile and
mature turtles (NQAA Fisheries Pacific Islands Regional Office. 2015b). Green sea
turtles are valued for their natural beauty, and several conservation groups are dedicated
to increasing sea turtle populations (Komoroskc et al.. 2011). Turtles can also be a source
of ecotourism (Campbell. 2002).
June 2018
14-27
DRAFT: Do Not Cite or Quote

-------
1	Final Ecosystem Services:
2	• Direct Human Uses:
3	• Cultural: Some native Hawaiians consider the green turtle to be a personal
4	or family deity, which should not be harmed. The green turtle is featured in
5	Hawaiian culture through depictions in petroglyphs (NQAA Fisheries Pacific
6	Islands Regional Office. 2015a).
14.7.4. White Ash
7	Scientific Name: Fraxinus americana.
8	• Family: Oleaceae (olive).
9	Symbolic Role: NA.
10	Federal or State Threatened or Endangered Species Listing Status: No special status.
11	Geographic Range/Distribution:
12	• Native Range: Most of eastern North America from southern Canada to northern
13	Florida and as far west as Minnesota (Griffith. 1991).
14	• Current Range: In addition to its native range, white ash is cultivated in other
15	areas, including Hawaii (Griffith. 1991).
16	Overlap with Class I Areas:
17	• http://www.epa.gov/visibilitv/maps.html.
18	• http://www.epa.gov/ttn/oarpg/t 1 /fir notices/classimp. gif.
19	Habitat: Usually found mixed with other hardwoods (not the dominant species) in
20	various topographic conditions. It can grow in a wide variety of soil types but does best
21	in soils with high nitrogen and calcium content, as well as moist and well-drained soils
22	(USD A. 1990a).
23	Primary Threats:
24	• Related to N or S Deposition: (Hernandez et al.. 2016).
25	• Other: White ash has experienced extensive dieback and mortality in its native
26	range since the 1920s, due to a variety of factors, including ash yellows (AshY
27	phytoplasma), canker fungi, viruses, drought (Hibbcn and Silverborg. 1978). and
28	more recently, invasions of the emerald ash borer [Agrilus planipennis; (Poland
29	and McCullough. 2006)1. White ash has also been found to be sensitive to ground
30	level ozone exposures (Chappelka et al.. 1988).
June 2018
14-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
Ecosystem Role and Function:
•	Nutrition: Samaras (seeds) are good forage for wood ducks, northern bobwhites,
purple finches, pine grosbeaks, fox squirrels, mice, and many other birds and
small mammals. It is browsed by white-tailed deer and cattle, and its bark is
occasionally used as food by beavers, porcupines, and rabbits (Griffith. 1991).
•	Cover: It provides cover to primary cavity nesters, such as red-headed,
red-bellied, and pileated woodpeckers, and to secondary nesters such as wood
ducks, owls, nuthatches, and gray squirrels (Griffith. 1991).
•	Other: As with other hardwood tree species in general, white ash trees provide a
number of other intermediate ecosystem services that are indirectly valued by
humans, including nutrient cycling, carbon sequestration, soil stabilization and
erosion control, water regulation and filtration, and air pollutant filtration
(Krieger. 2001).
Final Ecosystem Services:
•	Direct Human Uses (Final Ecosystem Services)
•	Raw material in wood products and related industries: White ash trees
provide a dense, shock resistant, and durable wood that is used in a variety of
products, including baseball bats, tool handles, furniture, cabinets, canoe
paddles, snowshoes, and railroad ties (USDA NRCS).
•	Energy/fuel source: Used for fuel wood (USDA NRCS).
•	Human health: Leaves from the white ash have reported beneficial uses,
including the relief of swelling and itching from mosquito bites and
prevention of snake bites (Griffith. 1991).
•	Cultural: A large number of ethnobotanical uses of white ash by Native
Americans has been documented. The Native American Ethnobotany
database maintained by University of Michigan (2015) references documents
citing 38 specific uses by 9 different tribes, including a wide variety of
medicinal uses (e.g., gastrointestinal, dermatological, and gynecological aids)
and uses as a material for snow gear, basketry, furniture, canoes, hunting, and
fishing.
•	Aesthetic: Provides attractive and vivid fall foliage, including combinations
of orange, yellow, maroon, and purple. For landscaping uses, it is also
considered to be a good and relatively quick-growing shade tree (University
of Kentucky. 2015).
•	Nonuse value: Survey-based economic studies have found evidence of
nonuse values among the general population for protecting white ash from air
pollution in the Adirondacks region (Banzhaf et al.. 2006).
14-29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
14.7.5.
Lace Lichen
Scientific Name: Ramalina menziesii (Ramalinaceae).
Federal or State Threatened or Endangered Species Listing Status: Listed as a
conservation species of least concern.
Geographic Range/Distribution:
•	Native Range: Lace lichen is native to the western coast of North America; in the
past it was found in the San Jacinto Mountains near Los Angeles and on the
surrounding coastal plain but now only grows at elevations above the smog layer
(Hastings Reserve. 2015).
•	Current Range: Lace lichen grows along the Pacific Coast of North America
from Alaska to Baja California. It grows at elevations up to 3,500 feet ("Werth and
Sork. 2008).
Overlap with Class I Areas:
•	http://www.epa.gov/visibilitv/maps.html.
•	http://www.epa.gov/ttn/oarpg/t 1 /fir notices/classimp. gif.
Habitat: Grows on tree species in coastal areas with enough moisture from fog and
strong winds (Egan and Holzman. 2003).
Primary Threats:
•	Related to N or S Deposition: (Hernandez et al.. 2016).
•	Other: Other threats to lace lichen include changes in climate. This species
requires a good amount of moisture and a temperate climate to survive; it is absent
from drier areas within its range (Egan and Holzman. 2003).
Ecosystem Role and Function: Lace lichen has a mutualistic relationship with its host
tree species. Three common host species in California are the valley oak, blue oak, and
coastal live oak (Werth and Sork. 2008). The lichen benefits from the microclimate and
structure provides by the host tree and it benefits the tree by capturing wind-borne
nutrients and moisture and depositing them back to the surrounding soil (Knops et al..
1996). This species is particularly useful in the Santa Lucia Mountains where nitrogen
limits plant growth (Hastings Reserve). Lace lichen is a food source for browsing animals
like deer and cows. It is also used as a nesting material for hummingbirds and orioles
(Hastings Reserve). Lace lichen has also been found to contribute to the global uptake of
sulfur dioxide, an airborne pollutant (Gries et al.. 1997).
Final Ecosystem Services:
•	Direct Human Uses:
• Aesthetic: Lace lichen has a unique morphology that adds to the natural
beauty of coastal woodlands along the Pacific Coast (Hastings Reserve.
2015).
June 2018
14-30
DRAFT: Do Not Cite or Quote

-------
1	• Education/research: Lace lichen is very sensitive to air pollution, including
2	exposure to nitric acid (Riddcll et al.. 2008) and can be used as an indicator
3	species to measure air quality (Gciscr et al.. 2010). New growth of lichen can
4	also indicate improved air quality conditions.
June 2018
14-31
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
APPENDIX 15. OTHER ECOLOGICAL EFFECTS OF
PARTICULATE MATTER
Evidence for effects of particulate matter (PM) on ecological receptors include direct
effects of airborne PM on radiative flux and both direct and indirect effects of deposited
particles. In addition to nitrogen (N) and sulfur (S) and their transformation products
(discussed in Appendix 3-Appendix 12). other PM components such as trace metals and
organics are deposited to ecosystems and may subsequently impact biota. Direct effects
include alteration of leaf processes from deposition of PM ("dust") to vegetative surfaces
(Appendix 15.4.1). Indirect effects encompass physiological responses associated with
uptake of PM components and alterations to ecosystem structure and function. Exposures
can occur directly to surfaces of vegetation (leaves, bark, twigs), by ingestion, or via soil
or water through uptake by roots or other biological tissues. Bioaccumulation and
biomagnification of PM components can lead to effects at higher trophic levels. Direct
and indirect effects of PM on vegetation, fauna, and the soil environment, and evidence
for effects at higher levels of biological organization (communities, ecosystems) from
PM deposition are evaluated in the context of what was known from previous PM
assessments (U.S. EPA. 2009a. 2004) and newly available studies.
15.1. Introduction
PM-associated components include N and S and their transformation products, trace
metals, organics, base cations, and salts. Base cations (especially Ca2+, Mg2+, and K+)
from atmospheric deposition can help ameliorate the effects of acidic deposition
associated with oxides of N and S, although under very high base cation deposition, plant
health can be adversely affected (U.S. EPA. 2009a). Particulate salt may be added to an
ecosystem from deicing salt (U.S. EPA. 2009a'). The focus of this appendix is on the PM
effects not associated with N and S components that are covered elsewhere in this ISA.
The primary non-N and S components covered in this appendix include trace metals
(Appendix 15.3.1) and organics (Appendix 15.3.2).
Effects of PM on ecological receptors can be both chemical and physical (U.S. EPA.
2009a. 2004). As described in the 2009 Integrated Science Assessment for Particulate
Matter (2009 PM ISA), particulates that elicit direct and indirect effects on ecological
receptors vary in terms of size, origin, and chemical composition. Particle composition is
attributed to ecological outcomes to a greater extent than particle size (Grantz et al..
2003). PM-associated metals and organics are linked to responses in biota; however, the
heterogeneous nature of PM composition and distribution coupled with variability
June 2018
15-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
inherent in natural environments confound assessment of the ecological effects of
particulates.
Ecological effects evaluated in the 2009 PM ISA included direct effects on metabolic
processes of plants, the contribution of PM to total metal loading which alters soil
biogeochemistry and microbiology, the effects on plant and animal growth and
reproduction, and the contribution to total organics loading resulting in trophic transfer.
Newly available information on the ecological effects of PM published since the 2009
PM ISA is summarized in the following sections along with key studies from previous
PM assessments (U.S. EPA. 2009a. 2004). Studies conducted in the U.S. are the focus of
this review, but research from other countries is included if it has advanced the study of
PM effects on biota. Such studies may include those providing documentation of
additional physiological effects of PM toxicity, new techniques for PM assessment, and
further characterization of effects on communities and ecosystems. Together with the
information available in the 2009 PM ISA, the body of evidence is sufficient to infer a
likely causal relationship between deposition of PM and a variety of effects on
individual organisms and ecosystems.
Direct Effects of Particulate Matter on Radiative Flux
The direct effects of particles suspended in the atmosphere on radiative flux and
subsequent effects on vegetation were described in previous PM assessments (U.S. EPA.
2009a. 2004). Briefly, increased atmospheric particulates and aerosols can alter radiative
flux by both radiation attenuation and by changing the efficiency of radiation interception
in the vegetation canopy by converting direct radiation to diffuse radiation (Hovt. 1978).
Diffuse radiation is more uniformly distributed throughout the vegetation canopy and can
reach leaves more evenly than direct radiation, increasing canopy photosynthetic
productivity. In a study reviewed in the 2009 PM ISA, decreased crop yield in China was
associated with reduction in solar radiation due to regional haze (Chamcidcs et al.. 1999).
More recent studies provide additional evidence of the role of PM in altering radiative
flux and plant response. In an assessment of aerosol effects on plant productivity in the
eastern U.S., increased canopy photosynthesis due to increases in diffuse radiation,
despite a simultaneous decrease in direct radiation due to light scattering by aerosols, was
observed on a regional scale (Matsui et al.. 2008). Surface downwelling solar radiation
was reduced by 14.9 W/m2 in 2000 and 16.0 W/m2 in 2001, while photosynthesis and
stomatal conductance increased simultaneously. In a recent study from Beijing, China,
the expression of plant proteins important to photosynthesis were shown to be directly
impacted by presence of aerosols and PM in the air column (Yan et al.. 2014). Expression
of plant ribulose 1,5-biphosphate carboxylase/oxygenase (RuBisCO), a key enzyme
15-2
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
involved in plant photosynthesis, increased in areas with strong diffuse solar radiation,
(identified by measurement of aerosol optical depth and particles with diameters of 0.1 to
1.0 |im). In the same areas, light-harvesting complex II protein and oxygen-evolving
enhancer protein decreased. These enzymes play a role in plant responses to direct solar
radiation when photosynthesis capacity is exceeded.
15.2. Particulate Matter Deposition to Ecosystems
Once airborne PM is deposited to aquatic or terrestrial systems, it can elicit additional
effects on ecosystem receptors. Large particles tend to be deposited near their sources,
such as along roadsides or next to mining, smelting, and other industrial operations, while
smaller particles can be transported long distances (Grantz et al.. 2003). Deposition to
ecosystems can be in the form of wet, dry, or occult deposition (U.S. EPA. 2009a; Grantz
et al.. 2003). Once deposited, PM-associated components may remain on biological
surfaces or be transferred between environmental compartments (e.g., water, soil,
sediment, biota).
More often, the chemical constituents drive the ecosystem response to PM, rather than
size class (U.S. EPA. 2009a; Grantz et al.. 2003). Exposure to a given mass concentration
of PM may lead to widely differing phytotoxic and other environmental outcomes
depending upon the particular mix of PM constituents involved. Deposition of the metal
and organic constituents of PM are discussed below; however, a rigorous assessment of
each chemical constituent (e.g., polyaromatic hydrocarbons [PAHs], mercury [Hg],
cadmium [Cd]) is not given. An overview of trace metals and organics and their
depositional effects in ecosystems was provided in the previous PM assessments (U.S.
EPA. 2009a. 2004) and only summarized below.
15.2.1. Metals
All but 10 of the 90 elements that comprise the inorganic fraction of the soil occur at
concentrations of <0.1% (1,000 jxg/g) and are termed "trace" elements or trace metals.
Trace metals with a density greater than 6 g/cm3, referred to as "heavy metals" (e.g., Cd,
copper [Cu], lead [Pb], chromium [Cr], Hg, nickel [Ni], zinc [Zn]), are of particular
interest because of their potential toxicity to plants and animals. For example, plant
toxicity to trace metals is most frequently associated with Cu, Ni, and Zn (U.S. EPA.
2004). Some metals such as Zn, manganese (Mn), Cu, and iron (Fe) play important
physiological roles in trace quantities; however, other metals such as Hg, Cd, Cr, and Pb
have no known biological function. Although some trace metals are essential for
June 2018
15-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
vegetative growth or animal health, they are all toxic in large quantities. Trace metals are
naturally found in small amounts in soils, groundwater, and vegetation. They may enter
the ecosystems as both fine and coarse particles. Heavy metals tend to be associated with
fine particles. Trace elements exist in the atmosphere in particulate form as metal oxides
(Ormrod. 1984). Aerosols containing trace elements derive predominantly from industrial
activities. Once deposited to biological surfaces, metals can be taken up by biota,
accumulate in tissues, and elicit toxic effects (Gall et al.. 2015). These responses are
highly variable across organisms. Only the bioavailable fraction is available for uptake by
biota. In plants, metal uptake is generally via soil to root transfer (McBridc et al.. 2013).
although recent studies reviewed in Appendix 15.4 provide evidence for foliar transfer of
metals following atmospheric deposition. The ecological effects of atmospherically
deposited Pb, a metal associated with PM, are included in the ISA for Pb (U.S. EPA.
2013b).
15.2.2. Organics
Organic compounds can be in the gas phase or in association with particles (Grantz et al..
2003). Organic compounds that may be associated with deposited PM include persistent
organic pollutants (POPs), pesticides, semivolatile organic compounds (sVOCs), PAHs,
and flame retardants, among others (U.S. EPA. 2009a). Dry deposition of organic
materials is often dominated by the course fraction, but fine PM may act as a carrier for
materials such as pesticides. PAHs are widespread in the environment. Of the
anthropogenically derived PAHs, 8 are considered carcinogenic and 16 have been
classified by the U.S. EPA as priority pollutants. In general, high molecular weight PAHs
are more toxic and have a greater tendency to be associated with PM (Mcsquita et al..
2014a). Derived from vehicular traffic and other sources, they are common air pollutants
in metropolitan areas. PAHs and other organics can be transferred to higher trophic
levels. Some atmospheric contaminants like POPs, polybrominated diphenyl ethers
(PBDEs), and other brominated flame retardants have been shown to accumulate in biota
in polar regions and at other remote locations where they are carried by long-range
atmospheric transport from lower latitudes (U.S. EPA. 2009a). The Western Airborne
Contaminants Assessment Project (WACAP) summarized in the 2009 PM ISA provided
evidence for airborne transfer and bioaccumulation of organics to remote national parks
and Class I areas (Landers et al.. 2008). and newer studies from this project have added to
this information (Pritz et al.. 2014; Landers et al.. 2010). Seven ecosystem compartments
(air, snow, water, sediments, lichens, conifer needles, and fish) were analyzed for a suite
of contaminants to determine where the pollutants were accumulating, identify ecological
indicators, and assess the source of the air masses most likely to have transported the
June 2018
15-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
contaminants to the parks. Overall, semivolatile organic compounds and metals were
detected throughout park ecosystems, and biomagnification of organics through the food
web indicated contaminants were present at multiple trophic levels.
As reviewed in previous PM assessments, vegetation itself is an important source of
hydrocarbon aerosols (U.S. EPA. 2009a. 2004). Terpenes, particularly a pinene, (3 pinene,
and limonene, released from tree foliage may react in the atmosphere to form submicron
particles. These naturally generated organic particles contribute significantly to the blue
haze aerosols formed naturally over forested areas (U.S. EPA. 2004; Geron et al.. 2000).
15.3. Effects of Particulate Matter on Vegetation
Particles deposited to foliar surfaces may lead to both structural and functional alterations
in plants. PM may physically obstruct processes associated with the leaf surface or be
taken up across the cuticle and into the plant tissues where it affects plant metabolic
activities. This section first considers direct impacts of PM deposition to vegetative
surfaces. Next, effects following uptake of PM components into plant tissue across leaf
surfaces or root uptake via soils are described. Uptake by plants can occur at the
soil/plant interface and at the air/plant interface (Krupa et al.. 2008). Once uptake occurs,
plant physiological responses may include decreased gas exchange, altered metabolism
and photosynthesis, altered pigment and mineral content, and altered enzyme activity
(Naidoo and Chirkoot. 2004). There is some evidence that metals reduce frost hardiness
and impair nutrition (Taulavuori et al.. 2005; Kim et al.. 2003).
15.3.1. Vegetative Surfaces
As described in the 2009 PM ISA, foliar surfaces are covered with a waxy cuticle layer
that helps reduce moisture loss and ultraviolet (UV) radiation stress (U.S. EPA. 2009a).
This epicuticular wax consists largely of long chain esters, polyesters, and paraffins,
which accumulate lipophilic compounds. Organic air contaminants in the particulate or
vapor phase can be adsorbed to, and accumulate in, the epicuticular wax of leaf surfaces.
Low solubility limits foliar uptake and direct heavy metal toxicity because trace metals
must be brought into solution before they can enter leaves or bark of vascular plants (U.S.
EPA. 2009a). Particles embedded in waxes may remain for extended periods of time and
remain associated with the leaf while other particles may be removed via precipitation,
wind, or leaf-fall. Particle capture efficiency varies by plant species, and the role of
vegetative surface characteristics in trapping PM, especially in polluted urban
environments, is extensively reviewed in the literature. PM trapping by vegetation
June 2018
15-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
depends on distance from source, leaf shape, canopy structure, leaf surface wettability,
presence of hairs, properties of the epidermal layer, and phyllotaxy (Popek et al.. 2015;
Popek et al.. 2013; Saebo et al.. 2012).
The 2009 PM ISA also reviewed studies that show effects of direct deposition of
particulates to aboveground plant organs. These effects include altered plant metabolism
and photosynthesis by the blocking of sunlight, obstruction of stomatal apertures,
increasing leaf temperature, and leaf surface injury (U.S. EPA. 2009a; Naidoo and
Chirkoot. 2004; Grantz et al.. 2003). Leaf surface pH can be altered by deposition of
alkaline particles, which can hydrolyze epicuticular surfaces (Grantz et al.. 2003). The
diverse composition of ambient PM and the effects of other air pollutants confound
characterization of the direct effects of PM on foliar surfaces (Grantz et al.. 2003).
A number of recent studies published since the 2009 PM ISA have examined the effects
of PM deposition to vegetative surfaces. For example, some have further characterized
the effects of particle size and composition on leaf surface processes. Following 60-day
experimental application of urban dust collected from roadsides in India to common
annual plant species (Abelmoschus esculentus, Celosia cristata, Coleus blumei,
Cyamopsis tetragonolobus, Gomphrena globosa, Impatiens balsamina, Ocimum sanctum,
Phaseolus vulgaris, Solanum melongena, Zinnia elegans), scanning electron micrographs
of leaf surfaces indicated that larger particles piled up around the stomata, while finer
particles clogged the stomatal openings (Rai et al.. 2010). In the applied dust, about 75%
was in the 2.5 to 10 |im size fraction, with course particles composing approximately
10% of PM and ultrafines comprising 15%. Experimental application of PM
(monometallic PM oxides: cadmium oxide [CdO], antimony trioxide | Sb:Ch |. and zinc
oxide [ZnO] and process PM enriched with Pb) to cabbage (Brassica oleracea) and
spinach (Spinacia oleracea) leaves resulted in a coverage rate of particles on the leaf
surface estimated at approximately 2% following washing (Xiong et al.. 2014). Most of
the particles were concentrated in stomatal apertures, with up to 12% of the area
occupied. The presence of PM in the stomata was correlated to particle size and solubility
of the associated metal oxides. Newer studies (Qguntimehin et al.. 2010; Qguntimehin et
al.. 2008) support observations of visible foliar injury reported in previous PM
assessments, with some suggesting that drought may affect the particle capture efficiency
of trees, possibly leading to decreased drought tolerance (Burkhardt and Parivar. 2015.
2014; Rasanen et al.. 2014).
15-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
15.3.2. Foliar Uptake of Particulate Matter
As reported in the 2009 PM ISA, fine PM has been shown to enter the leaf through the
stomata and penetrate into the mesophyll layers where it alters leaf chemistry and
physiology (Da Silva et al.. 2006; Naidoo and Chirkoot. 2004). Organic compounds can
be deposited as particles on the leaves or be taken up through the cuticle or stomata in the
gas phase, although these pathways are not well characterized for most trace organics
(Oguntimchin et al.. 2010). As reviewed in the 2009 PM ISA, for lipophilic POPs, such
as polychlorinated dibenzodioxins (PCDDs) and polychlorinated biphenyls (PCBs), the
air/plant response route generally dominates (Lee et al.. 2003; Thomas et al.. 1998). but
uptake through aboveground plant tissue also occurs. The pathways depend on the
chemical and its physical properties, such as lipophilicity, water solubility, vapor
pressure, and Henry's law constant. Environmental conditions can also be important,
including temperature and organic content of soil, plant species, and the foliar surface
area and lipid content.
(Eichcrt et al.. 2008) demonstrated hydrophilic particle penetration into the substomatal
cavity of faba bean (Vicia faba) leaves. Since the 2009 PM ISA, development and
refinement of techniques such as micro x-ray fluorescence, scanning electron microscopy
coupled with energy dispersive x-ray microanalysis, and secondary ion mass
spectrometry have been applied toward characterizing particulate transfer from foliar
surfaces to the interior of the plant (Schrcck et al.. 2014; Burkhardt et al.. 2012; Schreck
et al.. 2012). Repeated deliquescence and efflorescence of soluble PM may have the
effect of slowly advancing solutes into the stomatal pore where they may become
bioavailable (Burkhardt et al.. 2012). Recently, stomatal uptake of aqueous solutions was
confirmed by environmental scanning electron microscopy (Burkhardt et al.. 2012).
Deposited aerosols on leaf surfaces may alter surface tension and hydrophobicity,
enabling stomatal liquid water transport and facilitating foliar uptake. Several studies
provide additional evidence of particles entering plant tissues through stomata openings
(Schreck et al.. 2012; Uzu et al.. 2010). Schreck et al. (2014) observed particles in
stomata and damage to guard cells from Pb-rich particles from a Pb recycling factory.
Other pathways of deposited metals to vegetative surfaces may include diffusion across
the cuticle (Schreck et al.. 2012; Uzu et al.. 2010). Nonstomatal uptake of atmospheric
Hg into plant leaves has recently been demonstrated in addition to stomatal pathways
(Stamenkovic and Gustin. 2009). Biogeochemical transformations were observed in
metal-rich PM at the leaf surface (Schreck et al.. 2012). The role of the phyllosphere
(i.e., the microbial communities on foliar surfaces) on the uptake and chemical
transformation of deposited particles and in promoting plant growth is an emerging area
of study; however, interactions associated with PM are not well characterized at this time
(Wevens et al.. 2015; Xiong et al.. 2014).
June 2018
15-7
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
15.3.3.
Particulate Matter Impacts on Gas Exchange Processes
In earlier assessments of ecological effects of PM, deposition of particles to vegetation
was known to interfere with gas exchange processes such as photosynthesis, respiration,
and transpiration (U.S. EPA. 2009a. 2004). Several new studies support observations
from previous PM reviews on the exchange of oxygen and carbon dioxide across the leaf
surface. Most studies involve the application of dust to leaves. A greenhouse study with
lettuce (Lactuca serriola) leaves showed decreased gas-exchange parameters (net
photosynthetic rate and stomatal conductance) and a change in transpiration rate
following PM (dust from vehicular traffic with a dominance of fine particles) application
to leaves (Pavli'k et al.. 2012). Fly ash dusting of 0.5 to 1.5 g/m2/day to rice reduced
photosynthesis, stomatal conductance, and transpiration (Raja et al.. 2014).
Photosynthesis, transpiration rate, and water use efficiency was decreased in peach
(Primus persica) leaves dusted with cement kiln dust, and to a lesser extent with soil dust
(Malctsika et al.. 2015). Exposure of the leaves of Prunus padus to PM in Poland resulted
in decreased photosynthesis and chlorophyll fluorescence, but similar effects on the
photosynthetic apparatus were not found in Prunus serotina (Popck et al.. 2017). In a
study of native and non-native plant species in Argentina, Gonzalez et al. (2014) reported
that gas-exchange parameters, including maximum assimilation rate, stomatal
conductance, and transpiration rate, were significantly decreased in most plant species by
ambient dust accumulation on leaf surfaces compared with those parameters in plants
from which deposited dust had been removed. Similarly, decreased photosynthesis rate,
increased stomatal resistance, and lower chlorophyll content was observed with increased
PM accumulation in plant species growing in the city center of Warsaw, Poland
compared to a moderately clean environment (Przvbvsz et al.. 2014). The photosynthetic
apparatus of the lichen species Evernia prunastri was significantly affected by dust
enriched by calcium (Ca), Fe, and titanium (Ti) near a quarry and cement factory in
Slovakia (Paoli et al.. 2015). In open-top chamber experiments, leaf net photosynthesis
rate decreased in maize (Zea mays) plants exposed to 50 ng Hg/m3 compared to plants
exposed to 2 ng Hg/m3 (Niu et al.. 2014). Photosynthetic rate and chlorophyll content
were significantly reduced in cherry tomatoes (Lycopersicon esculentum) misted with
fluoranthene for 30 days to represent atmospheric wet deposition to the leaf surface
(Oguntimehin et al.. 2010).
15.3.4. Plant Physiology
In the 2009 PM ISA, particulates were associated with phytotoxic responses, including
induction of phytochelatins, alteration of pigments, and changes in mineral content and
enzyme activity (U.S. EPA. 2009a). Newly available studies support these findings.
June 2018
15-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
Chlorophyll content as well as specific leaf area and relative water content of selected
tree species was significantly reduced by the greater dust load associated with a polluted
site in India compared to a low pollution location (Chaturvcdi ct al.. 2013). Decreased
leaf chlorophyll and increased phenol content was also observed in peach leaves with
accumulated cement or soil dust (Malctsika et al.. 2015). In lettuce leaves, PM
application caused a depletion in amino acids that are metabolized during photosynthesis
as well as an increase in proline (Pavli'k et al.. 2012). Proline and malondialdehyde
concentrations were elevated in maize leaves treated with Hg (Niu et al.. 2014). Recently,
the use of foliar fatty acid composition as an indicator of foliar (and root) metal uptake
was demonstrated with lettuce leaves exposed to PM from smelter emissions (Schrecket
al.. 2013).
15.3.5. Uptake of Particulate Matter by Plants from Soils
Plants can also uptake PM that has been deposited to soils. As reviewed in the 2009 PM,
ISA uptake from soils varies by the PM component and plant species (U.S. EPA. 2009a).
There was some evidence that shallow-rooted plant species are most likely to take up
metals from the soil (Martin and Coughtrev. 1981). The ability of plants to take up
contaminants from soil and water has been applied toward environmental cleanup efforts,
a process known as phytoremediation (Hooda. 2007; Padmavathiamma and Li. 2007;
Clemens. 2006). Plants that are hyperaccumulators of metals are especially useful in
phytoremediation efforts (Prasad and DeOliveira. 2003).
Plants respond to high concentrations of metals in soil through a variety of mechanisms,
and plant species differ substantially in their response to heavy metal exposure. As
reviewed in the 2009 PM ISA, mechanisms of metal tolerance included exclusion,
excretion, genetics (Yang et al.. 2005; Patra et al.. 2004). mycorrhizal interactions (Gohre
and Paszkowski. 2006). storage capability and accumulation (Clemens. 2006). various
cellular detoxification mechanisms (Gratao et al.. 2005; Hall. 2002). and chelation with
phytochelatins (Memon and Schroder. 2009).
Since the 2009 PM ISA, additional studies have focused on the use of vegetation to
remediate sites with both organic and inorganic contaminants and the use of mycorrhizal
fungi and bacteria to promote phytoremediation (Appendix 15.5.4). In addition,
soil-bound PAHs associated with soil organic matter, historically thought to be generally
not easily available for root uptake, have been shown in recent studies to be readily
adsorbed to root surfaces, with some evidence showing translocation to the shoots
(Desalme et al.. 2013).
June 2018
15-9
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
15.3.6.
Effects on Plant Growth and Reproduction
PM effects on plant growth reviewed in the 2009 PM ISA include reduced vigor and
impaired root development (U.S. EPA. 2009a). In general, plant biomass is negatively
correlated with organics (Dcsalmc et al.. 2013) and metal concentrations (Audet and
Charcst. 2007). Inhibition of aboveground biomass was observed in lettuce leaves with
traffic-associated PM applied directly to vegetation (Pavlik et al.. 2012). Increased trace
metal concentration along with the decrease in biomass was measured in the treated
leaves. Decreased primary root elongation and shoot and root rates were observed in
tomatoes (Solarium lycopersicum) grown for 18 days in particulate matter with a nominal
mean aerodynamic diameter less than or equal to 10 ^m (PMio) collected from an urban
background site in Italy (Darcsta et al.. 2014). Decreased chlorophyll a and increased
carotenoids in the exposed plants were noted along with increased reactive oxygen
species production in roots from the PMio substrate. Phenanthrene, a common PAH in
air, applied to soils via simulated atmospheric deposition in an exposure chamber was
shown to significantly decrease root length and shoot biomass of leek (Allium porrum)
seedlings (Dcsalmc et al.. 2012). In a similar study, shoot biomass of red clover
(Trifolium pratense) exposed to atmospheric phenanthrene decreased by around 30%
while ryegrass (Loliumperenne) was unaffected (Dcsalmc et al.. 201 la). No effects on
root biomass were observed in either plant. With phenanthrene exposure to red clover,
more carbon (C) was retained in leaves with decreased C allocation to stems and roots
(Desalme et al.. 201 lb).
Daily application of urban dust to common annual plant species for 60 days showed
reduction in plant growth, number of leaves, and leaf area (Rai et al.. 2010). Growth and
yield of rice was significantly influenced at higher rates of fly ash deposition due to
increased heat load and reduced intracellular carbon dioxide (CO2) concentration (Raja et
al.. 2014). Following dusting of the plants by fly ash at 0.5, 1.0, and 1.5 g/m2/day, a
significant decrease in grain yield of 12.3, 15.7, and 20.2%, respectively, was observed.
Farahat et al. (2016) studied the effects of dust deposition from a rock quarry on
old-growth eastern hemlock (Tsuga canadensis) in Quebec, Canada. Mean radial growth
declined 43% after the construction of the quarry compared with growth rates before
construction. Major changes in the hemlock occurred 3-16 years after the quarry was
established, indicating that the effect on growth may be indirect through the alteration of
soil pH in this forest In a study from Beijing, China, application of local road dust to
Sophora japonica seedlings significantly affected growth characteristics such as leaf N,
shoot biomass, root-shoot ratio, photosynthesis, and total chlorophyll (Bao et al.. 2016;
Bao et al.. 2015).
June 2018
15-10
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
Recently, the effect of air pollutants on the timing and budburst of vegetation has been
assessed in several studies (Jochncr et al.. 2015; Kozlov et al.. 2007). The timing of full
flowering of hazel (Corylus avellana) was found to be significantly related to PM in
ambient air in urban areas of Munich, Germany (Jochncr et al.. 2015).
These effects on phenological onset were also observed with ozone (O3), nitrogen dioxide
(NO2), and nitrate (NO3) in hazel and additional tree species in the same study. In a field
and greenhouse study in Russia with mountain birch (Betula pubescens), no impacts of
air pollution on phenology were observed between trees surveyed in proximity to a Ni-Cu
smelter and trees at further distance from the emission source (Kozlov et al.. 2007).
Some studies have examined the effects of PM on plant reproduction. Some potential
effects demonstrated were those in flowering phenology and timing (Jochner et al.. 2015;
Rai et al.. 2010; Honour et al.. 2009). Jaconis et al. (2017) examined the effect of PM on
stigmatic clogging in Cichorium intybus on road sides in Cincinnati, OH. The authors
reported no relationship between PM levels and pollen germination across all of the roads
studied.
15.4. Effects of Particulate Matter on the Soil Environment
As described in the 2009 PM ISA, the soil is one of the most dynamic environments of
biological interaction in nature (U.S. EPA. 2009a). The upper soil layers where deposited
particles accumulate are typically active sites of litter decomposition and plant root
uptake. Processes within the rhizosphere, the soil around plant roots that mineral nutrients
must pass through, may be affected by PM components. Soil-associated microbial
communities of bacteria and fiingae such as actinomycetes and basidiomycetes,
respectively, break down organic matter for nutrient cycling and make elements available
for plant uptake. Soil mycorrhiza, (fungi that colonize plant roots) form a symbiosis to
provide nutrients in exchange for carbon from the plant. The effects of PM-associated
organics and metals on the soil environment is largely dictated by bioavailability to soil
microflora and plant roots.
15.4.1. Bioavailability in Soils
Soils are heterogeneous and the effects of PM deposition and subsequent bioavailability
of PM components depend on soil characteristics (e.g., pH, organic content).
PM-associated organics partition between the soil and the atmosphere based on soil
properties, such as organic matter content, moisture, porosity, texture, and structure, as
June 2018
15-11
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
well as the physiochemical properties of the pollutant, including vapor pressure and water
solubility (U.S. EPA. 2009a'). These compounds may be taken up by roots or be
associated with organic matter (Fismcs et al.. 2002). For heavy metals, accumulation in
soil is influenced by a variety of soil characteristics, including pH, Fe and aluminum
oxide content, amount of clay and organic material, and cation exchange capacity [CEC;
(Hernandez et al.. 2003)1. Bioavailability depends on metal speciation, soil pH, and
degree of binding to dissolved organic matter [DOM; (Sauve. 2001)1.
In the 2009 PM ISA, several models, isotopic studies, and sequential extraction methods
for determining metal bioavailability in soils were reported (Feng et al.. 2005; Collins et
al.. 2003; Shan et al.. 2003). However, actual measurement of biological effects is
generally regarded as the preferred criterion to assess bioavailability (Almas et al.. 2004).
Since the 2009 PM ISA, Biomet biosensors have been used to assess bioavailability of
metals in contaminated topsoils and subsoils (Almendras et al.. 2009). Biomet biosensors
use genetically engineered bacteria that give off light when exposed to heavy metals that
are bioavailable. Almendras et al. (2009) also used sequential extraction to determine
solubility. At three sites near a smelter in Chile, they found that Cu and Zn had the
greatest solubility, arsenic (As) had less solubility, and Pb and Fe were the least soluble
of those heavy metals. Further, as was determined to be the most bioavailable, while Cu
and Zn were less bioavailable. The authors were unable to determine bioavailability of Pb
using the same technique.
The mobility of heavy metals, an indicator of potential bioavailability, was determined in
surface soil using sequential extraction (Svendsen et al.. 2011). In five sites at decreasing
distances from a smelter in Norway, Cd was determined to be most mobile and Cu the
least mobile (Svendsen et al.. 2011). Zn and Pb were moderately mobile. These results
were in contrast to the results of (Almendras et al.. 2009).
15.4.2. Soil Nutrient Cycling
In previous PM reviews, toxicity of metals, especially Zn, Cd, and Cu, to soil microflora
was found to reduce decomposition processes in soils and interfere with nutrient cycling
(U.S. EPA. 2009a). In previous PM reviews, changes to microbial enzymatic activity, soil
basal respiration rate, and soil microbial biomass were all associated with increased metal
content in soils. Studies reviewed in the 2009 PM ISA report increased leaf litter
accumulation near point sources.
Since the 2009 PM ISA, many studies showing the deleterious effects of PM on nutrient
cycling have been published. The effect of PM on the microbial coefficient of soils
contaminated by atmospheric emissions from an active mining and smelting complex was
June 2018
15-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
determined in one study. The microbial coefficient, which is the ratio of microbial
biomass carbon to total organic carbon, was negatively correlated with metal
concentrations (Shukurov et al.. 2014). The microbial coefficient gives a measure of
substrate bioavailability.
A recent study demonstrated the role of atmospheric deposition of particulate heavy
metals in inhibiting C cycling. In this study, a cropping system was exposed to
atmospheric deposition while a similar cropping system was designed to exclude it by
using a greenhouse. Total organic carbon and water soluble organic carbon increased in
the exposed plots and decreased in the greenhouse plots; slow decomposition in soils
polluted with heavy metals could have caused the increased levels of total organic carbon
and water soluble organic carbon observed in the open plots (Pandev and Pandev. 2009).
New studies show decreases in enzyme activity and varying sensitivity of enzymes to
metal pollution. Soil dehydrogenase activity was diminished in polluted soils compared
to control soils (Boiarczuk and Kieliszewska-Rokicka. 2010). High soil dehydrogenase
activity indicates a living microbial community; additionally, dehydrogenase activity
increases with microbial biomass C, organic matter levels, and basal respiration
(Boiarczuk and Kieliszewska-Rokicka. 2010). In another study, enzyme activities of
alkaline phosphatase and fluorescein diacetate were greater in greenhouse treatments than
in open treatments exposed to deposition over the study period, showing that the potential
for hydrolysis was elevated in the greenhouse treatments (Pandev and Pandev. 2009). As
a result, organic C was mineralized to carbon dioxide and released from the soil more
quickly than it was in the treatments exposed to deposition. Qu et al. (2011) found that
enzyme activity was higher at sites farther from an active mining operation than those
close to it. However, urease activity was more affected by heavy metals than
dehydrogenase and phosphatase activity. Dehydrogenase contributes to many oxidative
activities, which degrade soil organic matter (Qu etal.. 2011). Phosphatase conducts
hydrolysis to convert organic phosphorus compounds into inorganic phosphorus
compounds, and urease converts urea into carbon dioxide and ammonia through
hydrolysis (Ou etal.. 2011).
A new study compared soil functional activity measured by the bait-lamina assay and by
the Biolog assay. Feeding activity was increased at the less polluted site compared to the
more polluted site (Boshoff et al.. 2014). Substrate utilization rate and richness of used
substrates were decreased in the more polluted site compared with the less polluted site
after 2 and 6 days. On the other hand, diversity of used substrates at all plots resembled
one another by the end of the experiment. Feeding activity, substrate utilization rate,
richness of used substrates, and diversity of used substrates all decreased as As, Cu, and
Pb concentrations increased. The Biolog assay was a better indicator of soil functional
15-13
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
activity than the bait-lamina assay (Boshoff ct al.. 2014). The bait-lamina assay indicates
the feeding activity of soil fauna; the Biolog assay measures the metabolic use pattern of
the culturable microbial community.
The effects of PM on microbial activity and biomass have been widely studied in the
literature since the release of the 2009 PM ISA. Additions of Cr and Zn to soils have been
found to reduce microbial respiration (Akcrblom et al.. 2007). The highest concentrations
of Pb and molybdenum (Mo) decreased respiration, as did higher levels of Ni and Cd. On
the other hand, low levels of Ni and Cd increased microbial respiration. In this study,
metals were added in low, medium, and high doses to soil from the humus layer of a
Swedish forest (Akcrblom et al.. 2007). Similarly, in another study, microbial activity
(for all metals except Cu) decreased with increasing metal concentrations, and a negative
correlation was also determined between metal concentrations and microbial biomass
(Anderson et al.. 2009a). Notably, microbial biomass levels were all similarly reduced
among contaminated sites compared to the control site, even though metal concentrations
among the contaminated sites varied. At the same time, microbial activity and biomass
increased with increasing concentrations of physicochemical variables (such as nitrate,
magnesium [Mg], and potassium [K]). Microbial activity, indicated by the amount of
carbon substrates used, was measured using the Biolog assay (Anderson et al.. 2009a).
Using similar methods, (Anderson et al.. 2009b) found that microbial activity and
biomass in polluted sites were reduced compared to activity and biomass in the control
site, prior to the experiment, which consisted of adding metals to soils from the sites.
Microbial activity and biomass were lower at all sites after metal addition (Anderson et
al.. 2009b). In Azarbad et al. (2013). basal respiration and substrate-induced respiration
increased with organic matter and pH and decreased as the toxicity index (an indicator of
heavy metal contamination) increased. The toxicity index was more important in
explaining effects of heavy metal contamination on substrate-induced respiration than
basal respiration. In the same study, microbial biomass, represented by total phospholipid
fatty acids (PLFAs), increased as toxicity index decreased (Azarbad et al.. 2013). Chodak
et al. (2013) found that microbial biomass and basal respiration were affected mainly by
environmental variables, such as total nitrogen and organic carbon, and heavy metals
impacted them much less. In another study, microbial biomass C and substrate-induced
respiration were lower in open treatments exposed to atmospheric deposition than in
greenhouse treatments as time progressed (Pandev and Pandev. 2009).
15.4.3. Soil Community Effects
At the time of the 2009 PM ISA, toxicity of heavy metals to soil-associated microbes
were well characterized in laboratory studies, but less was known about the relative
June 2018
15-14
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
sensitivity of fungi, bacteria, and actinomycetes and community-level changes. New
techniques described in the 2009 PM ISA, such as the use of phospholipid and
nucleic-acid biomarkers (Jovnt et al.. 2006). have enabled researchers to better
characterize microbial community composition and biodiversity. In a recent study,
tolerance of microbial communities to heavy metal PM has been shown to vary with
ecosystem type. Culturable bacterial communities in contaminated meadow soils
experienced greater tolerance to pollution, but the culturable communities in
contaminated forest soils did not (Stefanow icz et al.. 2009). The authors suggested that
similarities between the degree of heavy metal contamination in control and polluted
forest soils, caused by increased pH in polluted soils, may have contributed to the lack of
tolerance of bacterial communities in forest soils because pH influences bioavailability of
metals. Additionally, the absence of increased tolerance in forest soils might have been
caused by a lack of variation in metal sensitivity of bacterial species.
The impact of metal-associated PM on the abundance of microbes has been studied in
recently published papers. Total numbers of silver birch (Betulapendula) seedling root
tips colonized by mycorrhizae decreased in seedlings grown in polluted soils relative to
those in control soil (Boiarczuk and Kieliszewska-Rokicka. 2010). In another study,
culturable bacteria were more abundant at the sites farther from an active mining
operation than at the sites close to it (Qu etal.. 2011). Lenart-Boron and Wolnv-Koladka
(2015) found no correlation between metal concentrations and abundance of microbial
groups (mesophilic bacteria, fungi, actinomycetes, and Azotobacter spp.). The effects of
PM from heavy metals on the metabolic quotient of soils that were contaminated by an
active mining and smelting complex in Uzbekistan have been studied recently in the
literature. Metabolic quotient, an indicator of the influence of environmental variables on
microbial communities, increased with increasing heavy metal concentrations (Shukurov
etal.. 2014).
Recently published studies show that the response of soil-associated microbial
community structure and function to heavy metal PM varies. (Akcrblom et al.. 2007)
found that most metals altered community structure in a similar way. Phospholipid fatty
acids (PLFAs) can indicate microbial community structure, and different PLFAs are
associated with different groups of microbes [e.g., fungi, bacteria; (Akcrblom et al..
2007)1. Relative abundance of PLFAs associated with Gram-positive bacteria had a
positive relationship with metal concentrations, but relative abundance of PLFAs
associated with Gram-negative bacteria had a negative relationship with metal
concentrations. The relative abundance of actinomycetes PLFAs had a positive
relationship with metal concentrations. The reaction of the relative abundance of certain
PLFAs found in Gram-negative bacteria varied depending on exposure to particular
metals, and the reaction differed most between Cr and Cd. Relative abundance of a fungal
15-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
PLFA increased with increasing Cr and Zn concentrations (Akcrblom et al.. 2007). Likar
and Regvar (2009) determined that ascomycetes fungi were more common in polluted
plots, and ectomycorrhizal ascomycetes mainly comprised one taxonomic group. On the
other hand, basidiomycetes fungi were more common in control plots and in plots that
were less contaminated, and ectomycorrhizal basidiomycetes were more diverse than
ectomycorrhizal ascomycetes, comprising several taxonomic groups. Dark septate
endophytes in the ascomycete group were common in polluted plots and comprised
several taxonomic groups (Likar and Regvar. 2009). The increased number of DNA
sequences from a particular dark septate endophyte genus (Phialophora) in plots that
contained greater concentrations of Cd and Pb suggests that this genus could help goat
willow (Salix caprea) take up more nutrients under metal stress (Likar and Regvar.
2009); however, the authors did not actually measure nutrient uptake of plants in this
study. Anderson et al. (2009b) found that before the experiment, which consisted of
adding metals to soils with varying levels of initial contamination by an abandoned
smelter in Anaconda, MT, the structure and function of the microbial community from
each site differed from one another, and after the experiment, the communities remained
structurally and functionally different.
In another study, diversity and evenness of soil fungi communities in forest litter and
those of arbuscular mycorrhizal fungi communities in forest litter were lower in a site
polluted by emissions from an active copper smelter than they were in the reference site;
community structure of soil fungi in the polluted site was not similar to that in the control
site (Mikrvukov et al.. 2015). Spatial autocorrelations existed in both the control and
polluted sites for soil fungi communities, but the spatial structure of the communities in
the two sites did not resemble one another. Arbuscular mycorrhizal fungi (AMF)
community structures at control and polluted sites did resemble one another. Spatial
autocorrelation only existed in the polluted sites for arbuscular mycorrhizal fungi
communities (Mikrvukov et al.. 2015). Spatial heterogeneity tended to be greater in
contaminated areas than in uncontaminated areas (Mikrvukov et al.. 2015). Ou et al.
(2011) found that microbial community diversity increased as distance from an active
mining operation increased. In a different study, ecophysiological index values at two
long-term contaminated sites were substantially high for oligotrophs and copiotrophs
(Margesin et al.. 2011). Both sites contained bacterial communities with high Shannon
diversity index values (a common index used to characterize species diversity), and the
metabolically active groups at both sites were Proteobacteria and Actinobacteria
(Margesin et al.. 2011).
Recent studies have shown the effects of environmental variables in comparison to the
effects of heavy metal associated PM. In a study by Anderson et al. (2009a). microbial
species richness was influenced by physicochemical characteristics (Mg, CEC, Ca,
15-16
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
ammonium [NH/]), but was not affected at all by metal concentrations; however,
community structure in polluted sites was altered by metal concentrations compared to
control sites. In a different study, toxicity index, which represents the degree of heavy
metal pollution, had a positive impact on functional diversity of fungi, and nutrient status
had a negative impact on functional diversity of fungi (Azarbad et al.. 2013). On the other
hand, toxicity index did not influence functional diversity of bacteria, but functional
diversity increased as acidity decreased. PLFAs that represent Gram-positive bacteria
increased with toxicity index, while the PLFA associated with fungi decreased as toxicity
index increased. The level of organic matter and pH were factors that influenced PLFAs
as well as degree of metal contamination (Azarbad et al.. 2013). Chodak et al. (2013)
determined that pH had the largest impact on bacterial community diversity, but heavy
metals also modulated diversity. Bacterial community structure was mainly affected by
pH. Although pH was the main factor controlling microbial community structure, heavy
metals also influenced microbial community structure (Chodak et al.. 2013).
In addition to soil microbial communities, the effects of metal-associated PM on other
soil fauna have been recently published. Shukurov et al. (2014) found that abundance of
nematodes was greater at the plots further from the emission source; similarly, the
abundance of nematodes, number of bacterivorous nematodes, and number of
omnivorous predator nematodes were negatively correlated with heavy metal
concentrations.
The harmful effects of organic PM on soil microbes have been published in recent
studies. AMF infectivity in the top soils was reduced compared to control soils when
exposed to phenanthrene for 2 weeks (Desalme et al.. 2012). In this study, phenanthrene
was pumped into an exposure chamber containing soils to simulate atmospheric
exposure. However, in another study using a similar experimental setup, mycorrhizal
symbiosis in red clover was similar in control and polluted chambers, and mycorrhizal
symbiosis in both treatments remained efficient (Desalme et al.. 201 la). In the same
study, Rhizobium nodule (bacteria) symbiosis in clover was lower in the upper layer of
polluted soil than in the upper layer of control soil (Desalme et al.. 201 la).
15.4.4. Soil Microbe Interactions with Plant Uptake of Particulate Matter
In the 2009 PM ISA, the role mycorrhiza in modulating toxicity of deposited metals was
reported. This role includes improving nutrient uptake and decreasing metal uptake
(Vogelmikus et al.. 2006; Nogueira et al.. 2004; Berthelsen et al.. 1995) in the plant by
acting as a sink for metals (Carvalho et al.. 2006; Berthelsen et al.. 1995). often
preventing the metals in the roots from allocating to shoots (Spares and Siqueira. 2008;
June 2018
15-17
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Zhang et al.. 2005; Kaldorf et al.. 1999). In contrast, other studies reviewed in the
2009 PM ISA showed that mycorrhizae may facilitate accumulation of metals in plants
and enhance the translocation of metals from the root to the shoot (Zimmcr et al.. 2009;
Citterio et al.. 2005; Vogel-Mikus et al.. 2005). There was limited evidence for bacteria
and mycorrhiza working together to improve plant tolerance to metals (Vivas et al.. 2006;
Vivas et al.. 2003).
Since the 2009 PM ISA, additional studies show mycorrhizae modulates metal uptake in
plants. Seedlings grown in polluted soil with mycorrhizae preferentially sequestered
heavy metals (Cu and Pb) to the roots instead of the shoots (Boiarczuk and Kieliszewska-
Rokicka. 2010). The ratio of mycorrhizal fine roots to nonmycorrhizal fine roots was
lower for silver birch seedlings grown in polluted soils (mixture and fully polluted soil
near a copper foundry in Poland) than for tree seedlings grown in control soil; however,
the mycorrhizae were still able to affect metal uptake by the plants.
Nonmycorrhizal microbes (e.g., bacteria, fungal hyphae, spores from saprobes) may also
decrease metal uptake by plants. In a greenhouse experiment using soils collected at
varying distances from abandoned smelters in the U.S., wavy hairgrass (Deschampsia
flexuosa) plants grown in soils with high amounts of contamination and treated with
nonmycorrhizal microbial wash accumulated less Zn in the shoots compared to plants
without the nonmycorrhizal microbial wash (Glassman and Casper. 2012). In contrast,
although there was an AMF treatment in this experiment, AMF did not impact the
amount of Zn in the shoots of plants.
AMF and nonmycorrhizal microbes such as plant growth promoting bacteria can
contribute to phytoremediation of polluted soils; much new research on these topics has
been published since the release of the previous ISA (Cabral et al.. 2015; Meier etal..
2012; Gamalero et al.. 2009).
15.4.5. Effects of Particulate Matter on Physical Properties of Soils
PM can affect physical properties of soils, such as bulk density, porosity, and water
holding capacity. Changes in these properties can decrease plant growth and yield. This
topic was not discussed in the previous PM ISA. However, new studies on the effects of
PM on physical properties of soils have been published since the release of the previous
ISA. Pandev and Pandev (2009) found that bulk density was elevated during the study
period in an experimental cropping system open to deposition. Porosity and water
holding capacity in the open system were diminished over the same time period. The
opposite was the case for the experimental cropping system closed to deposition. The
authors suggested that the influx of deposition of particulate matter during the study
June 2018
15-18
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
period caused a rise in bulk density, and the elevated bulk density caused the decrease in
porosity and water holding capacity (Pandev and Pandev. 2009).
15.5. Effects of Particulate Matter on Fauna
Toxicity of PM to fauna varies depending upon the PM composition, concentration,
bioaccessibility and bioavailability, biological species sensitivity, and organism lifestage.
The bioavailability or bioaccessibility of PM-associated trace metals and organics is
dependent upon the physical, chemical, and biological conditions under which an
organism is exposed at a particular geographic location. Organisms may have
detoxification mechanisms that increase tolerance to stressors. Pathways of PM exposure
to aquatic and terrestrial organisms include ingestion, absorption, and trophic transfer.
Several types of studies have been used to assess the effects of PM on terrestrial and
aquatic organisms, including laboratory bioassays and field studies.
15.5.1. Laboratory Bioassays
The use of ecotoxicity assays to screen and assess PM effects on biota has expanded
considerably since the 2009 PM ISA. All but one of the new studies used PM from
outside of the U.S., so the pollutant mix may not be representative of exposures occurring
in the U.S. These studies are reported below. However, several caveats must be noted in
correlating effects in these studies to natural environments. Extracts of PM from air filters
may not be representative of exposure routes and conditions in aquatic and terrestrial
habitats (Kovats et al.. 2012). Bioavailability of PM in the natural environment is affected
by many factors not represented in bioassay techniques. Furthermore, because these
studies include a mix of PM constituents, it may be difficult to identify the active
component(s). Results from toxicity assays using PM extracts (such as dichloromethane,
methanol, or water extracts) should be interpreted with caution because additional
compounds might remain unextracted, underestimating or overestimating the whole PM
toxicity to aquatic organisms (Mesquita et al.. 2014a; Kovats et al.. 2012).
Recently the Vibrio fischeri bioluminescence inhibition test (Microtox®) has been applied
as an ecotoxicological screening tool for PM (Roig etal.. 2013; Kovats et al.. 2012). In
this bioassay, the rate of inhibition of light emission from the bacteria is measured after
exposure to a pollutant. In aqueous PM extracts from air filters from urban, rural, and
industrial areas of Catalonia, Spain, no correlation was observed between inhibition of
light emission in the in vitro test and PMio concentration (Roig et al.. 2013). Metals
(except Cr) and PCDD/F congeners in the samples were significantly correlated to the
June 2018
15-19
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
bioassay results, indicating that PM composition, rather than concentration, explains the
experimental observations.
Toxicity of PM extracts to aquatic biota have been assessed in several studies published
since the 2009 PM ISA. Acute (24-hour) toxicity of ambient PM with a nominal mean
aerodynamic diameter less than or equal to 2.5 |im (PM2 5) extracted from air filters in
Atlanta, GA to the freshwater rotifer Brachionus calyciflorus was assessed by (Vermaet
al.. 2013). The toxicity of methanol extracts was much higher than that of water extracts,
and there were no conclusive associations of rotifer toxicity with PM chemical
composition. In zebrafish (Danio rerio) embryos exposed to organic extracts of PM
samples from coal-burning particles in a semirural area in Italy, cytochrome P4501A
gene expression was significantly induced (Olivarcs et al.. 2013). In another zebrafish
study, urban PM had adverse developmental effects (spinal deformations, malformation
of swim bladder, etc.) on embryos (Mcsquita et al.. 2014b). These observations correlated
with PAH content of the PM samples. In a comparison of PM in rural and urban air
particles collected in Spain, induction of AhR signaling pathway correlated with PAH
concentrations in all locations (Mesquita et al.. 2015). The greatest dioxin-like activity
and embryotoxicity was observed in the finest PM fraction (<0.5 |im). During the winter,
maximal assay activity occurred in the rural samples due to biomass-burning emissions.
A series of in vivo studies on the effects of PM2 5 exposure on the nematode
Caenorhabditis elegans indicated a suite of responses in this organism. Zhao et al.
(2014b) observed altered development, lifespan, reproduction, locomotion, and
defecation behavior associated with long-term exposure to PM2 5 collected from a
traffic-dense area in Beijing, China. PM components included S, Cd, Pb, Zn, Cu, and
PAHs. The insulin signaling pathway was identified as a possible molecular target of the
traffic-related PM2 5 in nematodes (Yang etal.. 2016; Yang et al.. 2015). In nematodes
exposed to PM2 5 in coal ash, altered functioning of neurons that control defecation
behavior and gene expression patterns required for control of oxidative stress were
reported, as were effects on development, reproduction, locomotion behavior, lifespan,
and susceptibility to metal toxicity (Wu et al.. 2017; Sun et al.. 2016; Sun et al.. 2015b).
Elemental analysis of the PM from coal combustion indicated Fe, Zn, and Pb were the
most common metals followed by As, Cd, Cr, Cu, and Ni. Fluoranthene, pyrene, and
12 additional PAHs were detected in the PM2 5 in the coal ash.
Earthworms often constitute a large percentage of soil animal biomass and are considered
to be relatively sensitive bioindicators of soil contamination by metals and organics.
Several studies in the 2009 PM ISA reported uptake of trace metals and organics by
earthworms (Hobbelen et al.. 2006; Parrish et al.. 2006). while fewer studies reported
responses due to atmospheric deposition of PM (Massicotte et al.. 2003). Earthworms can
15-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
also modify PM bioavailability and bioaccessibility in soils through bioturbation. For
example, the presence of earthworms has been shown to significantly increase
soil-to-plant transfer of metals via modifications to soil and increased bioaccessibility of
metals to roots (Lcvcquc ct al.. 2014).
In a recent study, earthworms (Eisenia andrei), directly exposed to powder from PMio
quartz filters (24.9 mg/g PMio and 14 jxg/g PAHs) in artificial soil for a range of final
concentrations from 15 to 30 jxg/g PMio, showed genotoxic effects (Vcrnile et al.. 2013).
DNA damage measured by comet assay was observed starting at 22.5 jxg/g PMio
(0.012 jxg/g PAHs) in samples collected from an urban site in Italy.
15.5.2. Wildlife as Biomonitors of Particulate Matter
The 2009 PM ISA reviewed several studies in which resident biota was used to
biomonitor urban air pollution including PM. For example, snails (Helix spp.) accumulate
trace metals and agrochemicals and can be used as effective biomonitors for urban air
pollution (Regoli et al.. 2006; Viard et al.. 2004; Beebv and Richmond. 2002). Biological
effects that have been demonstrated include oxidative stress, growth inhibition,
impairment of reproduction, and induction of metallothioneins, which are involved in
metal detoxification (Regoli et al.. 2006; Gomot-De Vaufleurv and Kerhoas. 2000).
Biomonitoring of aquatic species was also reported. For example, Coelho et al. (2006)
investigated Hg concentrations in Scrobiculariaplana, a long-lived, deposit-feeding
bivalve in southern Europe. The use of sentinel species to detect the effects of complex
mixtures of air pollutants is of particular value because the chemical constituents are
difficult to characterize, exhibit varying bioavailability, and are subject to various
synergistic effects.
New biomonitoring studies of PM effects on biota include vertebrate and invertebrate
organisms. Use of the land snail Helix aspersa as a biomonitor for PAHs associated with
atmospheric particles in Porto Alegre, Brazil indicated, in general, that higher
genotoxicity in this indicator organism was associated with PM size fractions 
-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
in excreta were always higher in the wild urban pigeon population, while methemoglobin
levels in blood showed significant differences from season to season, but values in the
indoor birds were significantly lower only in summer and autumn (Sicolo et al.. 2009).
The authors chose this model because pigeons have a small habitat and high metabolic
rate, and the main route of exposure to urban air is through inhalation and intake of
contaminated food particles. Assay results were compared with measurements of
contaminants in the urban air (carbon monoxide [CO], PMio, NO2, O3, SO2, benzene
[CeHe], PAHs in PM2 5), and a positive correlation with PAH and protoporphyrin was
observed along with a positive correlation between O3 and DNA damage index (Sicolo et
al.. 2010; Sicolo et al.. 2009). Feathers, lung, liver, and kidney tissues were sampled from
pigeons in two urban areas in the U.S. (Glendora, CA and Midland, TX) to evaluate Hg
exposure (Cizdziel et al.. 2013); however, no significant differences were observed
between the two locations. In another study in two atmospherically polluted sites in North
Texas, Hg was found in the blood and feathers of eastern bluebird (Sialis sialis), Carolina
wren (Thryothorus ludovicianus), wood duck (Aix sponsa), great egret (Ardea alba), and
great blue heron (Ardea herodias) (Schulw itz et al.. 2015). Levels reported in these birds
may be high enough to affect fitness (Schulwitz et al.. 2015).
Several species have been evaluated as indicators of metal pollution from a long-term
monitoring site near a former Cu-Ni smelter in Haijavalta, Finland. Metal concentrations
in bird feces were quantified at the site to measure exposure via atmospheric deposition;
however, measured decreases in the excrement did not directly reflect emission patterns
(Bcrglund et al.. 2015). The authors suggest this is due to uptake of metals from food
items in contact with contaminated soils. Plasma carotenoid levels measured in the birds
on the site were not directly correlated to metal pollution or nesting survival (Ecva et al..
2012). Leg deformities of oribatid mites were also evaluated as a biomonitor but
determined to be an unreliable indicator of heavy metals in soil at the site (Ecva and
Penttinen. 2009).
15.5.3. Biomagnification
Biomagnification of PM-associated metals and organics was reviewed in the
2009 PM ISA and are summarized here. The basic understanding of the process of
biomagnification has not appreciably changed since the previous assessment. As reported
in the 2009 PM ISA, biomagnification is the progressive accumulation of chemicals with
increasing trophic level (Leblanc. 1995). At the time of the 2009 PM ISA,
biomagnification had been demonstrated for a few trace metals in terrestrial and aquatic
systems (U.S. EPA. 2009a). Plant uptake is often the first step for a metals to enter higher
levels of the food web. Consumers of vegetation may often receive heavy loading of
June 2018
15-22
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
metals from their diets. Metals may also bioaccumulate in some species, and tissue
concentrations are magnified at the higher trophic levels. Organic Hg is the most likely
metal to biomagnify, in part because organisms can efficiently assimilate methylmercury
and are slow to eliminate it lYCroteau et al.. 2005; Reinfelder et al.. 1998); Appendix 121.
There is also evidence that the trace metals Cd, Pb, Zn, Cu, and selenium (Se)
biomagnify.
Biomagnification of organics has been extensively documented in aquatic and terrestrial
ecosystems, and these compounds are detected in biota at remote locations due to
long-range atmospheric transport processes (U.S. EPA. 2009a'). Organics, such as PAHs,
can be transferred to higher trophic levels, including fish, and this transfer can be
mediated by aquatic invertebrates in the fish diet (U.S. EPA. 2009a'). In high mountain
lakes, atmospheric inputs dominate as sources of contamination. In addition, such lakes
tend to have relatively simple food webs. In a study reviewed in the 2009 PM ISA,
(Vives et al.. 2005) investigated PAH content of brown trout (Salmo trutta) and their food
items. Total PAH concentrations tended to be highest in organisms that occupy littoral
habitats, and lowest in pelagic organisms.
The study of trophic transfer and biomagnification is limited by the difficulty in
discriminating food webs and the uncertainty associated with assigning trophic position
to individual species (Croteau et al.. 2005). Use of stable isotopes can help establish
linkages. However, it is difficult to determine the extent to which biomagnification
occurs in a given ecosystem without thoroughly investigating physiological biodynamics,
habitat, food web structure, and trophic position of relevant species. Thus, developing an
understanding of ecosystem complexity is necessary to determine which species are at
greatest risk from toxic metal exposure (Croteau et al.. 2005).
15.6. Effects of Particulate Matter on Ecological Communities and
Ecosystems
Evidence for PM effects on higher levels of biological organization is primarily from
field studies conducted in proximity to point sources, such as smelters, mining, and other
industrial sources. These areas where PM deposition decreases from the source have been
assessed through ecological gradient studies. Several long-term monitoring studies have
been conducted on ecosystem responses to atmospheric deposition of metals, including
the Haijavalta smelter in Finland where metal-rich dust emissions have decreased by 99%
during the 23-year study period (Eeva and Lehikoinen. 2015). Additional evidence for
PM effects on ecological communities include studies from urban areas and model
microecosystems.
June 2018
15-23
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
15.6.1.
Gradient Effects near Smelters
The 2009 PM ISA reviewed multiple studies conducted near a Cu-Ni smelter in
Haijavalta, Finland. These studies documented effects on multiple species operating at
different trophic levels (Kiikkila. 2003; Helmisaari et al.. 1999; Malkonen et al.. 1999;
Pennanen et al.. 1996). The species composition of vegetation, insects, birds, and soil
microbiota changed, and tree growth was reduced 4 km from the smelter. Within 1 km,
very few organisms were documented (Kiikkila. 2003).
Since the 2009 PM ISA, additional studies have been published from the long-term
monitoring of the ecological communities near the Cu-Ni smelter site in Haijavalta,
Finland. Several studies in passerine birds (pied flycatcher [Ficedula hypoleuca\ and
great tit [Parus major]) show some biological recovery and a reduction in metal
concentration with decreased atmospheric emissions over time, even though metal levels
in tissues remain high in birds near the smelter, likely due to contamination of soils and
food items (Bcrglund et al.. 2012; Berglund et al.. 2011). Similarly, tissue levels in pied
flycatchers remained high near a smelter in Northern Sweden even after a 93 to 99%
reduction in metal emissions (Berglund and Nvholm. 2011). With decreased metal
pollution at Haijavalta, breeding parameters in pied flycatchers and great tits have
improved since the 1990s. However, clutch size and fledgling number remain below
those of birds from a reference area (Eeva and Lehikoinen. 2015; Eeva et al.. 2009).
Interruption of egg laying was assessed in passerine birds at the Haijavalta Smelter site
and was attributed to both cold weather and pollution (Eeva and Lehikoinen. 2010).
These laying gaps occurred with greater frequency, and most commonly at the beginning
of the laying sequence, close to the smelter. The authors suggested that heavy metals at
the site interfere with Ca availability and metabolism during the breeding season when
extra Ca is needed for egg shell formation and growth.
Eeva et al. (2010) sampled land snail shells from pied flycatcher nests along the
Haijavalta smelter pollution gradient to assess effects of air pollution on shell mass,
abundance, and diversity of land snail communities. The snails represent an important
source of Ca for the birds, especially for egg shell formation and development of
nestlings. Shell size decreased near the smelter, indicating snail growth was also
impacted at the most highly polluted areas. The highest diversity, largest size, and
greatest abundance of snail populations were observed in the moderately polluted areas
compared to snails in the vicinity of the smelter and in remote unpolluted areas.
Recent literature continues to show a gradient of response in vegetation to trace metal
deposition from smelting operations including evidence from the U.S. The forest
surrounding two historical Zn smelters in Palmerton, PA has been used as a study site to
assess ecological response and recovery following historic emissions. A portion of this
June 2018
15-24
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
forest includes the Appalachian National Scenic Trail, and a barren area of approximately
600 to 800 ha remains where Zn, Mn, and Cd were once emitted (Bever et al.. 2013).
Operation at the smelter ceased in 1980, but effects from the contaminated soils persist.
In a greenhouse study using seedlings of native tree species, as little as 10% of soil
collected in proximity to the smelter site and mixed with reference soil caused reduced
growth or mortality of the tree seedings (Bever et al.. 2013). Red maple (Acer rubrum)
was most sensitive, followed by gray birch (Betulapopulifolia), northern red oak
(Quercus rubra), chestnut oak (Quercus prinus), and eastern white pine (Pinus strobus).
In a field survey relating Zn soil concentrations to forest response, canopy closure and
shrub cover was decreased to half at approximately 2,060 mg/kg Zn, while tree-seedling
density was reduced by 80% at 1,080 mg/kg Zn (Bever etal.. 2011). Vegetative
communities near the Nikel smelter in Russia near the border with Finland and Norway
showed varied response to deposition of heavy metals (Mvking et al.. 2009). Epiphytic
lichens were most affected followed by decreased abundance and species richness of
lichens and bryophytes closer to the smelter. No effects on tree crown condition or
growth were observed.
15.6.2. Urban Environments
Resident biota in urban areas are affected by PM along with other stressors associated
with the built environment such as increased temperature due to heat island effects
(Pickett et al.. 2011). McDonnell et al. (1997) observed differences in urban forest
structure and function compared to rural and suburban areas of New York. Oak forests
near the city were characterized by reduced soil fungal and invertebrate populations,
elevated soil metals (presumably from air pollution), and lower quality of leaf litter. The
removal of PM by urban vegetation to mitigate air pollution has been extensively
reviewed, but few of these studies consider ecological impacts of PM (Escobedo et al..
2011; Pataki etal.. 2011).
15.6.3. Aquatic Ecosystems
In the 2009 PM ISA, PM components in aquatic ecosystems were primarily considered in
the context of aquatic food web transfer of organic contaminants (U.S. EPA. 2009a). In
another study reviewed in the 2009 PM ISA, bioassay procedures with green algae were
used to provide an initial screening of ambient PM toxicity from two urban/industrial and
one rural site near Lake Michigan (Sheeslev et al.. 2004). Results suggested that toxicity
was not related to the total mass of PM in the extract but to the chemical components of
the PM. In a study of the effects of contaminated tunnel wash water runoff in Norway,
June 2018
15-25
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
growth of sea trout (Salmo trutta) was significantly reduced in the stream receiving the
wastewater (Meland et al.. 2010). The contaminated wash water had elevated
concentrations of traffic-related contaminants including metals, PAHs, and road salt from
combustion and highway runoff. Fractionation of water samples indicated some of the
metals and PAHs were highly associated with particles. Size ratios of detected PAHs
suggested that sources included tire wear and incomplete combustion.
15.6.4. Experimental Microecosystems
Since the 2009 PM ISA, several studies have employed a microecosystem to assess
effects of air pollutants at the community level of biological organization to see how
microbial communities living on terrestrial mosses (bryophytes) respond to atmospheric
deposition of PM components (metals, PAHs, NO2). In a series of bioindicator studies
conducted in France, species-specific responses to PM components and effects on total
biomass were observed in the associated microbial communities (Meyer et al.. 2010a;
Mever et al.. 2010b). Primary producers, decomposers, and predators responded
differently to the pollutants (Mever et al.. 2010a; Mever et al.. 2010b). Testate amoebae
are especially sensitive to changes in atmospheric pollution in these microsystems and
were proposed as a biomonitoring tool (Mever et al.. 2012). Total testate amoebae
abundance and the abundance of five species of testate amoebae decreased with increased
deposition of the PAH phenanthrene in the experimental microecosystem (Mever et al..
2013).
15.7. Summary of Ecological Effects of Particulate Matter
Since publication of the 2009 PM ISA, new literature builds upon the existing knowledge
of ecological effects associated with PM components, especially metals and organics. In
some instances, new techniques have enabled further characterization of the mechanisms
of PM on soil processes, vegetation, and effects on fauna. New studies provide additional
evidence for community-level responses to PM deposition, especially in soil microbial
communities. However, uncertainties remain due to the difficulty in quantifying
relationships between ambient concentrations of PM and ecosystem response. Overall,
the body of evidence is sufficient to infer a likely causal relationship between
deposition of PM and a variety of effects on individual organisms and ecosystems,
based on information from the previous review and new findings in this review.
In regard to direct effects of PM on radiative flux, a newly available research method
links changes in expression of proteins involved in photosynthesis to changes in radiation
June 2018
15-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
associated with aerosols and PM. Although this method has not been widely applied, it
may represent an important way to study mechanistic changes to photosynthesis in
response to more diffuse radiation resulting from PM in the air column.
In general, new studies on PM deposition to vegetation support findings in previous PM
reviews on altered photosynthesis, transpiration, and reduced growth. Additional
characterization of PM effects at the leaf surface since the 2009 PM ISA has led to a
greater understanding of PM foliar uptake. Alterations in leaf fatty acid composition are
associated with metals transferred to plant tissues from PM deposition on foliar surfaces
(Schreck et al.. 2013).
Several studies published since the 2009 PM ISA show PM effects on soil physical
properties and nutrient cycling. Previous findings in the PM ISA of changes to microbial
respiration and biomass are further supported by new studies. Microbial community
responses to PM vary in tolerance to heavy metals and organics.
In fauna, results from ecotoxicity assays with PM extracts using bacteria, rotifers,
nematodes, zebrafish, and earthworms support findings in the 2009 PM ISA that toxicity
is not related to the total mass of PM in the extract, but to the chemical components of the
PM. In nematodes exposed to PM from air filters, the insulin-signaling pathway was
identified as a possible molecular target. Use of wildlife as PM biomonitors has been
expanded to new taxa since the last PM review. Several studies in invertebrates and birds
report physiological responses to air pollutants, including PM.
For ecosystem-level effects, a gradient of response with increasing distance from PM
source was reported in the 2009 PM ISA. Newly available studies from long-term
ecological monitoring sites provide limited evidence for recovery in areas such as those
around former smelters due to the continued presence of metals in soils after operations
ceased. A novel experimental microecosystem using microbial communities living in
terrestrial mosses indicates that PM deposition alters responses of primary producers,
decomposers, and predators.
15-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
APPENDIX 16. CASE STUDIES
This appendix includes six case studies which are meant to identify the ecological effects
of nitrogen (N) and sulfur (S) with a specific focus on national parks, other protected
areas, and areas with long-term research. The locations of these case studies were chosen
because they are areas for which a substantial amount of published work on ecological
response to N and/or S deposition is available. The locations include the northeastern
U.S., Adirondack State Park, southeastern Appalachia, Tampa Bay, Rocky Mountain
National Park, and southern California. These case studies identify current acidification
and nutrient status, as well as empirical and modeled critical loads. A recent model of
reactive N deposition to Federal Class I areas suggests that Nr deposition exceeds the
most sensitive critical loads in many of these areas [Figure 16-1. from Lee et al. (2016)1.
The scientific characterizations of ecological responses to N and S in these case studies
set the scientific foundation for further analysis of risk and exposure.
0.00	2.25	4.50	6.75	9.00 [kg N/ha/yr]
Notes: color indicates magnitude of the reactive N exceedance. The size of Class I areas is not represented. Air quality model grid
cells containing Class I areas are shown as colored regardless of the fraction of grid cell area the Class I area covers. Bold line
divides Western and Eastern U.S.
Source: Figure 9 from Lee et al. (2016).
Figure 16-1 CL exceedance in Class I areas.
June 2018
16-1
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
16.1.
Northeastern U.S. Case Study: Acadia National Park,
Hubbard Brook Experimental Forest, and Bear Brook
Watershed
16.1.1. Background
This case study is meant to identify effects of nitrogen (N) and sulfur (S) in the
northeastern U.S., with a specific focus on national parks and areas with long-term
research data. This case study identifies current acidification and nutrient status and
empirical and modeled critical loads (CLs). The 2008 NOx-SOx ISA included a case
study of acidification in the Adirondack region of New York [Section 3.2.2.4 of 2008
ISA (U.S. EPA. 2008aYI. This case study is considered as a supplement to that earlier
case study. Further information about the Adirondack region can be found in Appendix 4
and Appendix 5.
16.1.1.1. Description of Case Study Region
Acadia National Park (ACAD) in coastal Maine, the Hubbard Brook Experimental Forest
(HBEF) in the White Mountains of New Hampshire, and the Bear Brook Watershed
(BBW) in southeastern Maine were chosen for this case study to represent northeastern
U.S. ecosystems known historically to be sensitive to acidification and nutrient
enrichment from atmospheric S and N deposition (Figure 16-2. Table 16-1). In this case
study, we highlight multiple decades of research and monitoring at these locations to
provide insights into the ecosystems' responses to S and N deposition and look for any
indications of recovery.
June 2018
16-2
DRAFT: Do Not Cite or Quote

-------
Northeast Case
Study Region
I Caw Study locations
hm Native American
Reservations
USA City Populations
•	1,000,000 plus
•	500.000 999,999
250,000 - 499,999
•	100,000 - 249,999
•	50,000 99,999
10,001 • 49,999
New*,.-
Hampshire
Acadia National
Park
Oc*sn
Figure 18-2 Locations of northeastern U.S. case study areas and nearby
human population centers.
Table 16-1 Selected characteristics of northeastern case study areas.
Case Study Area
Elevation
Geology
Dominant Vegetation
Focus
ACAD
0 to 466 m
Granitic
Red spruce, mixed forest
National Park
HBEF
183 to 1,002 m
Metamorphic
Northern hardwood
forest, mixed forest,
spruce-fir forest
Experimental forest, site of
diverse biogeochemical research
and monitoring
BBW
172 to 450 m
Metamorphic
Northern hardwood forest
Site of watershed manipulation
experiment
ACAD = Acadia National Park; BBW = Bear Brook Watershed; m = meters; HBEF = Hubbard Brook Experimental Forest.
16.1.1.1.1. Acadia National Park
1	Acadia National Park is located at the interface between northern coniferous forests and
2	temperate deciduous woods and hosts plant species from two distinct regions
June 2018
16-3
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
(http://www.acadiacentennial2016.org/visit-acadia/acadias-treasures/'). ACAD
encompasses 45,000 acres on two islands and a mainland peninsula on Maine's coast. It
receives about 140 cm/year of rainfall (Nielsen and Kahl. 2007). Geologically, the park
consists of granite mountains with surrounding sedimentary and metamorphic rock. The
soils have low buffering capacity, and steep slopes at high elevations make soil and
runoff susceptible to acidification. The park's ecosystems include alpine heath, stunted
growth woodlands, jack pine forests that include pitch pine, rocky woodlands of black
spruce and heaths, mature spruce, and fir. Red spruce and sugar maple are the
predominant tree species.
Eighty freshwater plant species have been documented in ACAD with another
12 semiaquatic shorelines species. Seven of the aquatic and semiaquatic plants are listed
or proposed for inclusion on Maine's List of Endangered and Threatened Species.
Another 30 of these species are considered "locally rare"
(https://www.nps.gov/acad/learn/nature/plants.htm').
Hubbard Brook Experimental Forest
Hubbard Brook Experimental Forest is located in the southern portion of the White
Mountain National Forest in central New Hampshire (Figure 16-3). The Atlantic Ocean is
about 116 km to the southeast. Most of the surrounding land is also within the National
Forest. The area is hilly and steep in places with coarse-textured acidic soils. The bedrock
is dominated by metamorphosed igneous and sedimentary rocks. Bedrock is covered by
about 2 m of glacial till.
Northern hardwood forests occupy lower elevation slopes with spruce-fir occurring on
upper elevation slopes. Annual average precipitation is about 140 cm (one-third to
one-quarter of which falls as snow; http://hubbardbrookfoundation.org/wp-
content/uploads/2010/12/long term effects.pdf).
16-4
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
f - ^, Ml. Cuahman
ak^.977 m
-—b A\ \ - J
Bird Trsneect Ares
mk
I
Mt. w
Klneo *
1015 m \
o-..
.o
/ \
H %Ą>*
Legend
°-. /
° \ 8.	^
to—•" *v/	N
o Rain gage
	Stream
	Gaged Watershed o
	HBEF Boundary
A
3 km
FS = Forest Service; HBEh = Hubbard Brook Experimental Forest; km = kilometer; m = meter; Mt = mount; Mtn = mountain;
Natl = national.
Notes: shown are the network of rain gages, experimental watersheds(1-9), and Mirror Lake.
Source: Campbell et al. (2007).
Figure 16-3 Site map of Hubbard Brook Experimental Forest in the White
Mountains of New Hampshire.
Glacier-deposited materials vary greatly in degree of sorting and grain size, with
thicknesses ranging from zero on ridgetops and in stream valleys to 50 m near Mirror
Lake. The principal soils are acidic (pH about 4.5 or less) and relatively infertile. The
20- to 200-mm thick forest floor allows rapid infiltration of water and insulates against
soil freezing before snow accumulates. Streams vary from small ephemeral channels to as
large as perennial 5th order. Up to about 60 to 80% of storm precipitation flows by
stream. HBEF is entirely forested, predominately with deciduous northern hardwoods
including sugar maple (Acer sctcchanim), beech (Fagus grandifolia), and yellow birch
(.Betula cillegheniensis). High-elevation conifers include red spruce (Picea nibens),
balsam fir (Abies balsamed), and white birch (Betula papyri ferct). Logging operations
ended around 1915 to 1917. The second-growth forest is about 80 to 90% hardwoods and
10 to 20% conifers. HBEF hosts more than 90 species of birds, snowshoe hare, moose,
June 2018
16-5
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
fox, black bear, beaver, and white-tailed deer are present
(http://www.hubbardbrook.org/overview/sitedescription.shtml').
HBEF ornithological studies have found that the abundance of birds decreased from more
than 200 individuals per 10 ha in the early 1970s to 70 to 100 per 10 ha during the period
from the early 1990s to the present. This decrease is unexplained (R.T. Holmes
publications—Bird Abundances—http://\\\\\\ .hubbardbrook.org/data/datasct.php'.)id=S 1
and http://hubbardbrookfoundation.org/wp-
content/uploads/2010/12/long term effects.pdf).
Bear Brook Watershed
The BBW hosts a long-term, gaged, forested, first-order paired stream watershed. It is
located about 40 km from the Atlantic Ocean on the southeast slope of Lead Mountain. It
has a total relief of 210 m and a maximum elevation of 450 m. Two nearly perennial, low
dissolved organic carbon (DOC), low acid neutralizing-capacity (ANC) streams (East and
West Bear Brook [EB and WB]) drain 10.3 and 11.0 ha contiguous watersheds,
respectively. BBW is covered predominantly by hardwoods (Fagus grandifolia, Acer
rubrum, Acer saccharum, Betula alleghaniensis, Betulapapyrifera, and Acer
pensylvaticum) and red spruce. Soils are coarse, loamy, mixed, frigid Typic Haplorthods
developed on till averaging 1 m in thickness. Bedrock is mainly quartzites and
metapelites with local granitic intrusions
(https://umaine.edu/bbwm/about-bbwm/site-description-2/').
West Bear Brook has received bimonthly additions of ammonium nitrate since November
1989 (-1,800 eq/ha/yr—a 300% increase over ambient N deposition at the beginning of
the study). East Bear Brook has received no additions and serves as a reference [see
(Norton et al.. 1999)1. BBW has been a research site for three decades with a focus on
increasing decadal-scale understanding about northern forested ecosystems' response to
chemical and physical change, including whole ecosystem acidification, nitrogen
enrichment, and climate change. BBW research has focused on: alterations to nitrogen
dynamics, base cation decline, carbon cycling, phosphorus controls on nitrogen cycling,
response to decreasing ambient sulfate deposition, sustained elevated experimental
sulfate deposition, and changes in forest growth
(https://umaine.edu/bbwm/research/environmental-research/').
16-6
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
16.1.1.2. Class I Areas
In an effort to preserve pristine atmospheric conditions, the Clean Air Act (42 USC 7470)
authorized Class I areas to protect air quality in national parks over 6,000 acres in size
and national wilderness areas over 5,000 acres.
Class I areas are subject to the "prevention of significant deterioration (PSD)" regulations
under the Clean Air Act. (42 USC 7470). PSD preconstruction permits are required for
new and modified existing air pollution sources, and air regulatory agencies are required
to notify federal land managers (FLMs) of any PSD permit applications for facilities
within 100 km of a Class I area.
ACAD is a PSD Class I Area. HBEF is not in a Class I area, but is near two New
Hampshire Class I areas, the Great Gulf and the Presidential Range—Dry River
Wilderness Areas. BBW is located southwest of the Moosehorn Wilderness Area in
Maine.
16.1.1.3. Regional Land Use and Land Cover
No large population centers are located near HBEF or BBW. Several small towns are
interspersed with ACAD lands, including Bar Harbor (pop. 5,235), Southwest Harbor
(pop. 1,765), and Trenton (pop. 1,481). The populations of these towns increase
substantially with tourists during summer. Bangor (pop. 32,000) lies about 40 miles north
from Mount Desert Island. All three study areas are mostly forested, as is much of the
Northeast region (Table 16-2. Figure 16-4).
June 2018
16-7
DRAFT: Do Not Cite or Quote

-------
Table 16-2 Land use/land cover for northeastern case study areas.


Area Covered (km2)


ACAD
HBEF
BBW
Developed and barren
83.82
2.96
0.00
Hardwood forest
76.84
159.53
6.64
Conifer forest
592.56
61.88
0.60
Mixed forest and shrubland
300.27
111.49
2.53
Meadow/herbaceous
26.33
0.50
0.00
Wetland
137.05
1.47
0.16
ACAD = Acadia National Park; BBW = Bear Brook Watershed; HBEF = Hubbard Brook Experimental Forest; km = kilometer.
Northeast Case Study
Region: Land Cover
! Cast Study Locations
Native American
Reservations
Developed Land [increasing intensity)
Barren land
I Deciduous forest
H Evergreen Forest
Mixed Forest
	; Shrub/Scrub
Grassland/Herbaceous
¦ Cultivated Crops
m, Water/Wetlands
0	20	40	GO	80
Acadia .National
Park
Atlantic Ocean
Figure 16-4
June 2018
Land cover in the Northeast case study region.
16-8	DRAFT: Do Not Cite or Quote

-------
16.1.1.4. Organization of This Case Study
1
2
3
4
5
6
7
8
9
10
Table 16-3 Literature cited by Northeast U.S. case study area.
Variable
Acadia
BBW
HBEF Northeast Regional
Summary of studies from NE case study for acidification
Base cation

Elvir et al. (2006)

BAI

SanClements et al.
(2010)

Mineralization

Fernandez et al.
(2003)

Stream dissolved
S042", pH, ANC, Al+

Norton et al. (2004)

Base cation, Ca, Mg

Norton etal. (2004)
Gbondo-Tuqbawa
and Driscoll (2003),
Gbondo-Tuabawa
and Driscoll (2002)
Summary of studies from NE case study for nutrient status
Aquatic N cycling

Mineau et al. (2014).
Simon et al. (2010)

Soil N cycling

Mineau et al. (2014)
Jefts et al. (2004)

Ca addition


Battles etal. (2014)
N addition

Bethers et al. (2009)
Hunt et al. (2008)

June 2018	16-9	DRAFT: Do Not Cite or Quote
Because this case study addresses the condition of three discrete locations in the
Northeast (ACAD, HBEF, and BBW), we developed Table 16-3 to summarize the
primary post-2000 research reported in the case study. In addition, we cite relevant
research in other areas of the northeastern region. Appendix 16.1.2 presents information
about N and S deposition in the Northeast. Discussions of critical load or dose-response
research (Appendix 16.1.3) are each organized around the three case study areas and the
Northeast region. Appendix 16.1.4 presents information available on long-term
ecological monitoring, and Appendix 16.1.5 contains information on the current status
and forecasts for recovery in the case study areas. Key research literature since January
2008 is highlighted in Table 16-4.

-------
Table 16-3 (Continued): Literature cited by northeast U.S. case study area.
Variable
Acadia
BBW
HBEF
Northeast Regional
N retention + multiple
disturbances


Aber et al. (2002)

N retention
climate/freezing


Groffman et al.
(2011). Judd etal.
(2011). Campbell et
al. (2010)

N retention and fire
Nelson etal. (2007)
Campbell et al.
(2004b)



N dep effect on
wetlands
Calhoun etal. (1994)



N loading to coastal
waters
Nielsen and Kahl
(2007)



N addition to lakes
Saros (2014)



Summary of studies from NE case study for terrestrial critical load or dose-response
Dose-response trees
Ellis etal. (2013)
Elvir et al. (2006). Elvir
et al. (2003)

Pardo et al. (2011c).
Thomas et al. (2010).
McNultv et al. (2005)
Does-response herbs



Pardo etal. (2011c)
Hurd etal. (1998)
Dose-response soil

SanClements et al.
(2010) Fernandez et
al. (2003)
Gbondo-Tuabawa
and Driscoll (2002)
Pardo etal. (2011c)
Dose-response
surface water


Gbondo-Tuabawa
and Driscoll (2002)

Modeling climate,
acid-base chemistry,
changes in plants

Phelan etal. (2016)
Phelan etal. (2016)
Campbell et al.
(2009)

Summary of studies from NE case study for aquatic critical load or dose-response
Macroinvertebrate
production



Chadwick and Hurvn
(2005)
Shift from deposition
to climate influence



Mitchell and Likens
(2011)
Flow volume and peak
flows



Kim etal. (2010)
NO3" concentration



Aber et al. (2003)
N limitation



Baron et al. (2011a)
June 2018
16-10
DRAFT: Do Not Cite or Quote

-------
Table 16-3 (Continued): Literature cited by northeast U.S. case study area.
Variable
Acadia
BBW
HBEF
Northeast Regional
NO3" leaching



Driscoll et al. (2003a)
DOC



SanClements et al.
(2012)
Climate effects on
pools, concentrations,
and fluxes of major
elements



Pourmokhtarian et al.
(2012)
PH



Ouimet et al. (2006),
Dupont et al. (2005),
Ouimet et al. (2001)
ANC



Sutherland et al. (2015),
Nierzwicki-Bauer et al.
(2010), Tominaqa et al.
(2010), Ouimet et al.
(2006), Dupont et al.
(2005), Ouimet et al.
(2001)
Al+ = aluminum; ANC = acid neutralizing capacity; BAI = basal area increment; BBW = Bear Brook Watershed; Ca = calcium;
DOC = dissolved organic carbon; HBEF = Hubbard Brook Experimental Forest; Mg = magnesium; N = nitrogen; NE = northeastern;
N03" = nitrate; S042" = sulfate.
June 2018
16-11
DRAFT: Do Not Cite or Quote

-------
Table 16-4 Key recent research literature focused on the case study region.
Publication
Focus
Baron et al. (2011a)
Regional patterns in N limitation
Battles et al. (2014)
Ca addition at HBEF
Bethers et al. (2009)
Effects of S and N on sugar maple foliar chemistry
CamDbell et al. (2009)
Influence of climate on biogeochemical cycling
Campbell et al. (2010)
Monitoring of climate conditions at HBEF
Cho et al. (2009)
Response of stream high-flow chemistry to Ca addition at HBEF
Ellis et al. (2013)
Empirical critical loads
Elviret al. (2010)
Tree growth and foliar chemistry at BBW
Fernandez and Norton (2010)
Overview of BBW study
Fernandez et al. (2010)
Responses to experimental acidification at BBW
Groffman et al. (2011)
Soil freezing at HBEF
Hunt et al. (2008)
N and P content of leaves
Judd et al. (2011)
Soil freezing at HBEF
Kim et al. (2010)
Stream flow at BBW
Laudon and Norton (2010)
Episodic stream chemistry at BBW
Lawrence et al. (2012)
Recovery of soil base status
Lona et al. (2009)
Sugar maple decline
Mineau et al. (2014)
Enzyme activity in soil and leaf litter
Mitchell and Likens (2011)
S budget at HBEF
Mitchell et al. (2011)
S mass balances for HBEF and BBW
BBW = Bear Brook Watershed; Ca = calcium; HBEF = Hubbard Brook Experimental Forest; N = nitrogen; P = phosphorus; S = sulfur.
June 2018
16-12
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
16.1.2. Deposition
Characteristics of nitrogen and sulfur deposition affecting the ACAD, HBEF, and BBW
in the study area are shown in Figure 16-5 and Figure 16-6. Figure 16-5 A and
Figure 16-6A show 3-year average total deposition of N and S for 2011-2013;
Figure 16-5B shows the partitioning between oxidized and reduced N; Figure 16-6B
shows the 25-year-long time series for wet deposition for NOs . NH4+, SO42 and H+
obtained at the NADP/NTN (National Atmospheric Deposition Program/National Trends
Network) monitoring site at HBEF (NH02). Surrounding areas in Maine and New
Hampshire and inserts showing the coterminous U.S. (CONUS) are shown to place the
depositional environment in context. See Appendix 2 A, Appendix 2.5 and Appendix 2JS
for more information on deposition in the U.S. Other maps showing the contributions of
individual species to dry and/or wet deposition are given in Appendix 2/7.
Data shown in the map Figure 16-5 and Figure 16-6A were obtained from the hybrid
modeling/data fusion product, TDEP (Total Deposition,
http://nadp.sws.uiuc.edu/committees/tdep/tdepmaps/ and described in Appendix 2.7).
However, the time series of wet deposition is taken directly from data on the NADP/NTN
(Figure 16-6B). This was done to track changes in deposition since the passage of the
Clean Air Act Amendment (CAAA) because the Community Multiscale Air Quality
(CMAQ) model simulations used in TDEP extend back only to 2000.
June 2018
16-13
DRAFT: Do Not Cite or Quote

-------
Bear
Brook
Acadia
N Deposition (kg-N/ha)
*> J» & & •
Hubbard
Brook
O
O
Monitor NH# 02
Monitor ME # 98
Monitor Locations
Northeast Study Area
A
B
ha = hectare; kg = kilogram; N = nitrogen.
Figure 16-5 Total nitrogen deposition (A) and percentage of oxidized nitrogen
deposition (B) for the Northeast case study area estimated by the
National Atmospheric Deposition Program Total Deposition
Science committee.
June 2018
16-14
DRAFT: Do Not Cite or Quote

-------
600
S00
400
ro
JZ
~o
*1 100
c
o
2 200
o
CL
V
Q
100
Annual Wet Deposition and 3-Year Moving Average at
Site NH02: 1990 - 2014
• 			
• N.

a	.. ~ V _
1988
1992
1996
2000	2004
Year
2008
2012
2016
B
H+ = hydrogen ion; ha = hectare; kg = kilogram; mol = mole; NH4+ = ammonium; N03 = nitrate; S = sulfur; S042 = sulfate; yr = year.
Figure 16-6 Total sulfur deposition (A) for the Northeast case study area
estimated by the National Atmospheric Deposition Program Total
Deposition Science committee. Time series of wet deposition (B)
from the National Atmospheric Deposition Program/National
Trends Network in the Hubbard Brook Experimental Forest, NH.
June 2018
16-15
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
Comparison of Figure 16-5 A and Figure 16-6A indicates that the general pattern of
deposition ofN and S is broadly similar; deposition tends to be higher near the coast and
decreases inland. Deposition at ACAD is an exception, with lower values along the coast
than elsewhere. As seen in Figure 16-5A. total deposition of N is relatively uniform
within the three portions of the study area. In addition, the area surrounding the three
sites shows a high degree of regional homogeneity but with higher values in New
Hampshire. Deposition of N is not as high as in many areas of the central U.S. or the rest
of the Northeast. Deposition of S is considerably lower than along the Ohio River Valley.
Figure 16-5B shows that the deposition of nitrogen is estimated to be mostly in oxidized
form throughout the entire domain. Although most of the area is subject to N deposition
in oxidized form with highest percentages surrounding ACAD and BBW, there are areas,
principally in central Maine, where N is deposited mostly in reduced form. In
Figure 16-6B. wet deposition of all species shows that downward trends in NO3 , NH4 .
S042 and H+ are consistently found over the past 25 years, although the rate of decrease
has been variable. In general, wet deposition typically exceeds dry deposition of N and S
in this case study area.
16.1.3. Critical Loads and Other Dose-Response Relationships
16.1.3.1. Terrestrial
This section presents post-2000 ACAD, HBEF, and BBW findings on the dose-response
relationships of N and S to terrestrial ecology, as well as the critical N and S loads for
maintaining ecosystem health. Additional relevant information for the northeastern region
is summarized. The section presents findings on both empirical research and modeling
analyses.
16.1.3.1.1. Empirical Studies
Post-2000 findings from empirical studies of terrestrial dose-response relationships and
critical loads are summarized in this section. Table 16-5 summarizes the body of
empirical and modeling research identified.
June 2018
16-16
DRAFT: Do Not Cite or Quote

-------
Table 16-5 Terrestrial empirical and modeling research on the response of
nitrogen and sulfur deposition for the northeastern U.S.



Deposition/






Addition



Variable
Species
Response
(kg N/ha/yr)
Years
Site
Reference
Primary
NA
PnET-BGC Model. Predicted
Not specified
1999-2099
HBEF
Campbell et
productivity

increase due to future longer



al. (2009)


growing season




Base cation
NA
PnET-BGC Model. Historical
Not specified
1850-1995
HBEF
Gbondo-
Ca and Mg

forest cutting had little impact



Tuabawa and
depletion

on exchangeable cation soil



Driscoll


pools



(2003)
Base cation
Sugar
American beech and red
N addition: 8.4
1989-2003
BBW
Elvir et al.

maple,
spruce had lower foliar Ca,
(wet + dry)


(2006)

American
Mg, Zn concentrations;
(NH4)2S04




beech,
nutrient imbalance may offset
addition:




red
potential photosynthesis
WB 25.2




spruce
benefits






Sugar maple had higher






photosynthesis rates; no






decrease in Ca, Mg, Zn






concentrations




Tree growth Northern +
	hardwood —
Mortality
ND
8.4 N
¦ (wet + dry)/
+25 (NH4)2S04
1989-2002 BBW
Pardo et al.
(2011c)
Foliar %N
Foliar %Ca
NO3" leaching
Cation loss
Soil C:N
N
mineralization
Nitrification
Soil
respiration
Microbial
biomass
Tree growth Red 0
	spruce —
Mortality
ND
8.4 N	1989-2002 BBW Pardo et al.
¦ (wet + dry)/+25	(2011c)
kg (NH4)2S04
June 2018
16-17
DRAFT: Do Not Cite or Quote

-------
Table 16-5 (Continued): Terrestrial empirical and modeling research on the
response of nitrogen and sulfur deposition for the
northeastern U.S.
Deposition/
Addition
Variable Species	Response	(kg N/ha/yr) Years	Site Reference
Foliar %N	+
Foliar %Ca
NO3 leaching	+
Cation loss	+
Soil C:N
N	+
mineralization
Nitrification	+
Soil	ND
respiration
Microbial	ND
biomass
Base cation Sugar
maple
Mineralization Not
specified
Little evidence of BC
depletion but confounded by
ice storm litter mineralization
Storm caused increased
litterfall, accelerated
mineralization, and obstructed
temporal trends in soil
chemistry (17 yr)
8.4 N	Not
(wet + dry)/ + 25 specified
kg (NH4)2S04
BBW SanClements
etal. (2010)
Fernandez et
al. (2003)
8.4	1989-1998 BBW
(wet + dry)/+25
(NH4)2S04
Soil base Not	ForSAFE-VEG model.	Not specified Not	BBW and Phelan et al
saturation specified Expected future climate	specified HBEF (2016)
change was simulated to
cause increase, especially at
BBW.
Soil acid-base Not	ForSAFE-VEG model.	Not specified Not	BBW and Phelan et al.
chemistry specified Simulated future climate had	specified HBEF (2016)
a lesser effect at HBEF than
BBW, likely due to the
overwhelming influences of
high S and N deposition.
June 2018
16-18
DRAFT: Do Not Cite or Quote

-------
Table 16-5 (Continued): Terrestrial empirical and modeling research on the
response of nitrogen and sulfur deposition for the
northeastern U.S.
Variable Species
Response
Deposition/
Addition
(kg N/ha/yr)
Years
Site
Reference
Plant	Not	ForSAFE-VEG model,
communities specified Climate futures predicted by
and N	the IPCC of increased
enrichment	temperature and precipitation
will change plant communities
and N enrichment,
counteracting the acidifying
impacts of S and N deposition
on soil acid-base chemistry.
Not specified Not	BBW and Phelan et al.
specified HBEF (2016)
NO3 leaching NA
PnET-BGC Model. Increase
due to enhanced net
mineralization and nitrification
Not specified 1999-2099 NE
Campbell et
al. (2009)
Mineral	NA	PnET-BGC Model. Slight Not specified 1999-2099 NE	Campbell et
weathering	decrease due to reduced	al. (2009)
simulated soil moisture
(negative effect) and
increased temperature
(positive effect)
NO3"	Not	DayCent-Chem model,
leaching, specified Simulated future high N
GHG	deposition under climate
sequestration,	scenarios had increased
ecosystem C	ecosystem GHG
pools	sequestration. High N
increased stream NO3" fluxes
and ecosystem C pools.
Simulated N 2001-2075 ACAD Hartman et al.
deposition 3.9 to	and	(2014)
9.6 kg/ha/yr for	HBEF
HBEF and 4 to
11.3 kg/ha/yr for
ACADr
BBW = Bear Brook Watershed; BC = base cation; C = carbon; Ca = calcium; ForSAFE-VEG = Soil Acidification in Forest
Ecosystems; ha = hectare; GHG = green house gas; HBEF = Hubbard Brook Experimental Forest; IPCC = Intergovernmental Panel
on Climate Change; kg = kilogram; Mg = magnesium; N = nitrogen; NA = not applicable; ND = no data; NE = northeastern;
(NH4)2S04 = ammonium sulfate; N03" = nitrate; PnET-BGC = Photosynthesis and Evapotranspiration-Biogeochemical; S = sulfur;
WB = West Bear Brook; yr = year; Zn = zinc.
16.1.3.1.1.1. Acadia National Park
1	Ellis et al. (2013) estimated the CL exceedance for nutrient-N deposition to protect the
2	most sensitive ecosystem receptors in 45 national parks, based on CL ranges compiled by
3	(Pardo et al.. 2011c). Ellis et al. (2013) estimated the N CL for ACAD in the range of
4	3-8 kg N/ha/yr to protect hardwood forests and a similar range to protect lichens.
5	Cleavitt et al. (2015) conducted surveys and used FIA assessments of lichen species in
6	the Northeast, with particular focus on the four Class I wilderness areas in the region
7	(Lye Brook, VT; Great Gulf, NH; Presidential Range—Dry River, NH; and Acadia
June 2018	16-19	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
National Park, ME). In general, cumulative N and S deposition over the course of the
modeled period (2000-2013) was a more powerful explanatory factor than annual N or S
load for the lichen data. This work determined a critical load of 4.3-5.7 kg total N
(oxidized + reduced N)/ha/yr based on lichen species richness, the abundance of sensitive
species (higher species richness of cyanolichens and fruticose lichens below CL), and
thallus condition. Cumulative S and N deposition were both equally powerful predictors
of thallus condition, but no S critical load was determined (Cleavitt et al.. 2015).
Pardo etal. (2011c) estimated that ambient N deposition in parts of ACAD was higher
than estimated empirical CL values. Thus, these data suggest the possibility of
exceedance of nutrient-N CL in ACAD (see Table 16-6).
Table 16-6 Empirical critical loads for nitrogen in Acadia National Park, by
receptor, from Pardo et al. (2011c).
Critical Load (kg/ha/yr)
NPS	N Deposition Mycorrhizal	Herbaceous	Nitrate
Unit	Ecoregion	(kg/ha/yr)	Fungi Lichen Plant Forest Leaching
Acadia Eastern Temperate	5.2	5 to 12 4 to 8	17.5	3 to 8	8
NP Forests
ha = hectare; kg = kilogram; N = nitrogen; NP = national park; NPS = National Park Service; yr = year.
Ambient N deposition reported by Pardo et al. (20110) is compared to the lowest critical load for a receptor to identify potential
exceedance. A critical load exceedance suggests that the receptor is at increased risk for harmful effects.
16.1.3.1.1.2. Hubbard Brook Experimental Forest
Gbondo-Tuabawa and Driscoll (2002) used the PnET-BGC model to simulate the
responses of soil and surface water chemistry at the reference watershed at HBEF to
future emissions controls. Model performance was assessed using the normalized mean
absolute error and the efficiency as objective statistical criteria. The focus was on
comparing simulated chemistry with observed values between 1980 and 1998. Results
showed good agreement for stream Ca and SOr . However, stream NOs and Al
concentrations and soil solution Ca:Al ratios were over-predicted after 1990. Simulation
results suggested that emissions controls required by the CAAA will likely not result in
substantial changes in the future in "critical indicators such as soil base saturation, soil
solution Ca:Al, or stream ANC and Al concentrations."
June 2018
16-20
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
16.1.3.1.1.3. Bear Brook Watershed
Elvir et al. (2006) investigated the effects of increased N deposition on photosynthesis
and foliar nutrient content of sugar maple, American beech, and red spruce at the BBW.
Trees in the treated WB watershed (an addition of 25 kg N/ha/yr as [NFLJ2SO4) had
higher foliar N concentrations than EB reference trees. American beech and red spruce
displayed significantly lower foliar concentrations of Ca, Mg, and Zn. Sugar maple did
not show decreases in these nutrients and was the only species to have significantly
higher photosynthetic rates in the treated watershed as compared with the reference. This
result suggested that nutrient imbalances in American beech and red spruce in the
treatment watershed may have offset any potential photosynthetic benefits to these
species from higher N availability (Elvir et al.. 2006).
During the first 7 years of treatment at WB, sugar maple basal area increment (BAI) was
higher in WB compared with EB. After 8 years of treatment, however, the initial higher
sugar maple growth rate in WB decreased (Elvir et al.. 2010).
In the WB catchments, the BAI of sugar maple was enhanced 13 to 104% by the addition
of 25 kg N/ha/yr as (NFL^SC^. The BAI of red spruce was not significantly affected
(Elvir et al.. 2003).
Fernandez et al. (2003) reported results of quantitative soil excavations in 1998 that
measured soil pools of exchangeable base cations after 9 years of treatment at WB. The
treated watershed had lower concentrations of exchangeable Ca and Mg in all soil
horizons, with greater nutrient cation depletion in the O-horizon as compared with the
mineral soil. Depletion was also greater in conifer, as opposed to hardwood, stands. The
importance of treatment-caused base cation depletion was reinforced by MAGIC model
simulations. Estimates of watershed-wide exchangeable Ca and Mg (66 and 27 kg/ha)
were roughly comparable to the total cumulative excess stream Ca and Mg export in WB
after 9 years of treatment (55 and 11 kg/ha, respectively; Figure 16-7).
June 2018
16-21
DRAFT: Do Not Cite or Quote

-------
5X
©
Qi
*
C3
3

Ca = calcium; ha = hectare; kg = kilogram; Mg = magnesium; yr = year.
Source: Fernandez et al. (2003).
Figure 16-7 Annual stream calcium and magnesium export (paired bars), and
cumulative excess export in West Bear Brook compared to East
Bear Brook (line), over the study period 1989-2000 at the Bear
Brook Watershed experiment.
16.1.3.1.1.4. Other Northeastern Regions
1	Pardo et al. (201 lc) compiled data on empirical CLs for protecting sensitive resources in
2	Level I ecoregions across the CONUS against nutrient enrichment effects caused by
3	atmospheric N deposition. Available data on empirical CL of nutrient-N in the Northeast
4	suggested that the lower end of estimates of the CL for resource protection was
5	3 kg N/ha/yr. Values at or above this level of deposition loading are considered to have
6	increased risk of harmful effects for mycorrhizal fungi, lichens, and forest vegetation
7	(Table 16-7). Keeping levels below this value also helps prevent NO, leaching into
8	drainage water.
June 2018
16-22
DRAFT: Do Not Cite or Quote

-------
Table 16-7 Critical loads of nutrient nitrogen for the Northern Forests ecoregion.
Critical load or
Ecosystem N deposition
Component (kg/ha/yr) Reliability	Response	Comment	Reference
Tree	>3	#	Decreased growth of None	Thomas et al.
red pine, and decrease	(2010)
survivorship of yellow
birch, scarlet and
chestnut oak, quaking
aspen, and basswood
Lichens
4 to 6
(#)
Community
composition shift
Application of model
developed for marine
West Coast forest to
northern forests
Geiser et al.
(2010)
Ectomychorrizal
fungi
5 to 7
#
Change in fungal
community structure
None
Lilleskov et al.
(2008)
Herbaceous
species cover
>7 and <21
#
Loss of prominent
species
Response observed in
low-level fertilization
experiment
Hurd et al.
(1998)
Northern
hardwood and
coniferous forest
8
##
Increased surface
water NO3" leaching
None
Aber et al.
(2003)
Tree growth and
mortality
>10 and <26
#
Decreased growth
and/or induced
mortality
Response observed in
low-level fertilization
experiment in
old-growth montane
red spruce
McNultv et al.
(2005)
Arbuscular
mycorrhizal fungi
<12
(#)
Biomass decline and
community
composition change
Observed along a
Michigan N gradient
Pardo et al.
(2011c)
ha = hectare; kg = kilogram; N = nitrogen; N03 = nitrate; yr = year.
Reliability rate: ## reliable—a number of published papers show comparable results; # fairly reliable—the results of some studies
are comparable; (#) expert judgment—few empirical data are available, Critical Load based on expert judgment of those
ecosystems.
Source: Pardo et al. (2011c).
1	Thomas et al. (2010) analyzed Forest Inventory Analysis (FIA) data in the Northeast to
2	determine tree growth in response to a gradient of atmospheric N deposition from about 3
3	to 11 kg N/ha/yr. Some tree species showed increased growth across the N input gradient
4	(yellow poplar [.Liriodendron tulipifera], black cherry [Prunus serotina], and white ash
5	[Fraxinus americana]). Some showed highest growth at intermediate levels of N
6	deposition (quaking aspen [Populus tremuloides\ and scarlet oak [Quercus coccinea]).
7	Red pine (Pinus resinosa) exhibited growth decline across the gradient of increasing N
June 2018	16-23	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
deposition (Thomas et al.. 2010). Thus, it appeared that N deposition at ambient levels
can have both positive and negative effects on tree growth, depending on species and
deposition level.
In a high-elevation red spruce-balsam fir (Abies balsamea) forest in the Northeast, N
fertilization over 14 years led to a decrease in live basal area (LBA) with increasing N
additions. In control plots, LB A increased by 9% over the course of the study, while LB A
decreased by 18 and 40% in plots treated with 15.7 kg N/ha/yr and 31.4 kg N/ha/yr,
respectively (McNultv et al.. 2005).
There is little information available regarding the effects of N deposition on herbaceous
plants within northern hardwood forests in the Northeast. However, Hurd et al. (1998)
reported the results of experimental studies that added N at two and four times the
ambient N deposition level at several sites in the Adirondack Mountains. Herbaceous
plant coverage decreased after 3 years of fertilization, largely in response to shading
caused by enhanced growth of ferns. Additionally, information has been recently
published by Simkin et al. (2016) and is discussed in Appendix 62 of this ISA.
Modeling Studies
Evidence from research in northeastern forests indicates that direct and indirect effects of
climate change will cause changes in N and other biogeochemical cycling by altering
plant physiology, forest productivity, and soil processes. Campbell et al. (2009) reviewed
these relationships (Figure 16-8) and applied the PnET-BGC model in a northern
hardwood forest ecosystem at HBEF to test assumptions about interactions regarding
climate change. Model results suggested an increase in primary productivity due to a
longer growing season in the future, an increase in NOs leaching due to enhanced net
mineralization and nitrification, and a slight decrease in mineral weathering due to
reduced simulated soil moisture (negative effect) and increased temperature (positive
effect).
16-24
DRAFT: Do Not Cite or Quote

-------
Temperature, Precipitation, ""-v
and Extremu Weather Hvents ¦'
Fire
Insects-pathogens
Invasive species
A forest composition
Feedbacks
Soil trace gas fluxes
Plant respiration
Deforestation
Transpiration
Evaporation
Albedo
Soil abiotic: process *":l ¦ l""" ;	Soil biotic processes
ion exchange Hedox reactions	Decomposition
Adsorption-deserption	Mineralization
Weathering Stream export	Immobilization
Solute transport ^	Respiration
Ecosystem Services
Water quality and quantity
Carbon sequestration
Economics (wood products,
tourism, maple sugar)
Recreation and aesthetics
NPP = net primary production.
Source: Campbell et al. (2009).
Figure 16-8 Conceptual diagram showing the direct and indirect effects of
changes in temperature and precipitation on biogeochemical
processes in forests and on the services forests provide. Also
shown are feedbacks that further influence climatological effects.
16.1.3.2. Aquatic
1	This section presents post-2000 research findings at ACAD, HBEF, and BBW on the
2	dose-response relationships of N and S to aquatic ecology, as well as the critical N and S
3	loads for maintaining ecosystem health. Additional relevant information for the Northeast
4	region is summarized. Both empirical and modeling studies of aquatic dose-response
5	relationships and critical loads are included in this section.
June 2018	16-25	DRAFT: Do Not Cite or Quote

-------
16.1.3.2.1. Empirical Studies
1	Post-2000 findings from empirical studies of aquatic dose-response relationships and
2	critical loads are summarized in this section. Table 16-8 compiles the body of empirical
3	research identified.
Table 16-8 Aquatic empirical research on the response of nitrogen and sulfur
deposition for the northeastern U.S.
Variable
Response
Deposition
Addition Years
Site
Reference
Fish
Shift in the upper
Not specified
0 1960S-2007
HBEF
Warren et al.
community
mainstem of HBEF



(2008)
including
from the presence of




brook trout
at least three fish




(Salvelinus
species to only brook




fori tin alis)
trout (Salvelinus





fontinalis)




S042"
W6 had a 32%
Not specified
0 1963-1994
HBEF
Driscoll et al.

decline in the annual



(2001b)

volume-weighted



Likens et al. (2001)

concentration




(-1.1 [jeq/L/yr)




S042" + NOs"
W6 had declines in
Not specified
0 1963-1994
HBEF
Driscoll et al.

stream



(2001b)

concentrations of





strong acids





(-1.9 [jeq/L/yr)




Sum of base
-1.6 [jeq/L/yr
Not specified
0 1963-1994
HBEF
Driscoll et al.
cations




(2001b)
PH
Small but significant
Not specified
0 1963-1994
HBEF
Driscoll et al.

increases In stream



(2001b)

pH, from 4.8 to 5.0




S042"
Biogeochemical
Not specified
0 1965-2008
HBEF
Mitchell and Likens

control of SO42"



(2011)

export from forested





watersheds has





shifted from





atmospheric S





deposition to climatic





factors that regulate





soil moisture.




pH and ANC Decrease	Not specified 1,800 eq/ha/yr 7 yr	BBW Norton et al. (2004)
(NH4)2SC>4 acidification
treatment
June 2018
16-26
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
Table 16-8 (Continued): Aquatic empirical research on the response of nitrogen
and sulfur deposition for the northeastern U.S.
Variable
Response
Deposition
Addition
Years
Site
Reference
Stream
dissolved
S042"
Decreased
Not specified
1,800 eq/ha/yr
(NH4)2S04
1989-2007
BBW
Norton etal. (2004)
Fatemi etal. (2012)
Base cations
Decreased
Not specified
1,800 eq/ha/yr
(NH4)2S04
1989-2007
BBW
Norton etal. (2004)
Fatemi etal. (2012)
Al+
Increased
Not specified
1,800 eq/ha/yr
(NH4)2S04
1989-2007
BBW
Norton etal. (2004)
Fatemi etal. (2012)
ANC
Regional surface
water ANC did not
change significantly in
New England during
the 1990s
Not specified
0
NA
NE
U.S. EPA (2003)
Evans and
Monteith (2001)
Release of S
from internal
storage pools
to drainage
water
Increased
Not specified
0
1965-2008
General
Mitchell and Likens
(2011)
Al+ = aluminum; ANC = acid neutralizing capacity; BBW = Bear Brook Watershed; ha = hectare; HBEF = Hubbard Brook
Experimental Forest; kg = kilogram; L = liter; |jeq = microequivalents; eq = equivalents; NE = northeastern;
(NH4)2S04 = ammonium sulfate; N03" = nitrate; S = sulfur; S042" = sulfate; W6 = Watershed 6; yr = year.
16.1.3.2.1.1. Acadia National Park
No empirical aquatic critical loads or dose-response studies have been identified in the
literature for the ACAD case study area since 2000.
16.1.3.2.1.2. Hubbard Brook Experimental Forest
The cycling of N, S, and other elements in the Northeast is closely linked with aspects of
climate. Many recent studies have explored some of those linkages, and many of the
empirical studies of aquatic dose-response relationships have included aspects of climate
change. The potential for a shift from atmospheric deposition to climatic regulation of
watershed S biogeochemistry at HBEF was examined by (Mitchell and Likens. 2011).
More than four decades of continuous long-term data collected from four watersheds
were used to evaluate S budgets. "Analyses focused on the role of changing water
availability in affecting the amount of S mobilized from internal watershed sources and
the resultant changes in the S budget discrepancies over time... The biogeochemical
June 2018
16-27
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
controls on annual SO42 export in stream water [appear to have] shifted from
atmospheric S deposition to climatic factors that affect soil moisture."
With declining inputs of atmospheric S, the higher outputs of SO42 in drainage waters
relative to precipitation inputs appear to be driven by the S stored in the soil. These
biogeochemical responses at the HBEF might be amplified with further climate change,
resulting in greater annual stream discharge and watershed wetness. Such climatic change
will potentially increase SO42 mobilization and hence may slow the recovery of both
aquatic and terrestrial ecosystems from acidification.
16.1.3.2.1.3.	Bear Brook Watershed
Chadwick and Hurvn (2005) assessed the effects of N additions on secondary stream
macroinvertebrate production in BBW. Production did not vary between streams
(reference and treatment) but was 35% higher for both streams in the second year of the
study. Results suggested that patterns of litter input and channel drying, rather than N
input, controlled secondary production in these intermittent streams by altering resource
availability and invertebrate community structure. It appeared that these variables
overrode the effects of N supply, perhaps because N is not limiting.
Ongoing climate change will likely affect a wide range of biogeochemical processes in
northeastern forested ecosystems. Of particular interest are the effects on flow volumes
and peak flows. Kim et al. (2010) analyzed a nearly two-decades-long stream flow record
at EB and a longer record for the Narraguagus River (a proximate watershed to the
BBW). The focus was on improving understanding of high-flow events that have a
disproportionate impact on biogeochemical processes and fluxes. A moving window
analysis to evaluate the changing flood potential over time suggested upward trends in
the occurrence of high-flow events during recent decades.
16.1.3.2.1.4.	Other Northeastern Regions
Considering trends across the Northeast, Aber et al. (2003) found that surface water NO3
concentrations exceeded 1 (j,eq/L mainly in northeastern watersheds receiving more than
about 9 to 13 kg N/ha/yr deposition. Above this range, mean NO3 export increased
linearly with increasing deposition at a rate of about 0.85 kg NO3 -N/ha/yr for every
1 kg N/ha/yr increase in deposition, although at higher rates of deposition there was
considerable variability in N retention among watersheds (Aber et al.. 2003).
N limitation across the Northeast was evaluated by Baron etal. (2011a) who estimated
that 34% of 4,361 New England lakes represented in the Eastern Lakes Survey were
June 2018
16-28
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
likely to be N limited, based on having dissolved inorganic N (DIN):total P (TP) ratio (by
weight) less than 4. Although eutrophication is a concern, there is no published evidence
that eutrophication is occurring to any appreciable extent in the lakes in ACAD.
Dissolved organic matter (DOM) is important to a range of processes critical to aquatic
ecosystem functioning, including providing a microbial food source, attenuating light,
buffering pH, binding metals, and controlling the cycling of C, N, Hg, and P. Shifts in the
quality of DOM may affect aquatic ecosystem functioning. A chemical signature of
terrestrial DOM was used by SanClements et al. (2012) to support the hypothesis that
increased dissolved organic carbon (DOC) concentrations in surface water in the
Northeast since about 1993 have been driven mainly by decreasing acidic deposition and
increasing solubility of soil organic matter. Fluorescence spectroscopy was used to
characterize the quality of DOM of archived samples from nine acid-sensitive lakes in
Maine. Decadal decreases in SOr in lake water were correlated with increased DOC
concentrations and a shift from microbial to terrestrially derived DOM during ecosystem
recovery from prior acidification.
Aquatic biota in the Northeast have been affected by acidification at virtually all levels of
the food web. Some species and some lifestages are especially sensitive. Decreases in
ANC and pH and increases in inorganic Al concentration contributed to declines in
species richness and abundance of zooplankton, macroinvertebrates, and fish as
documented in the Adirondacks (Sutherland et al.. 2015; Nierzwicki-Bauer et al.. 2010;
Keller and Gunn. 1995; Schindler et al.. 1985). Although some species are favored by
increased acidity, the overall species richness typically decreases as surface water acidity
increases.
Modeling Studies
Post-2000 findings from modeling studies of aquatic dose-response relationships and
critical loads are summarized in this section. Table 16-9 compiles the body of research
identified.
16-29
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Table 16-9 Critical and target load and exceedance modeling studies in the
northeastern U.S.
Reference
Location Model
Focus
Dupont et al. (2005)
NE region SSWC
CLs for acidity to protect aquatic biota


to pH 6.0
SanClements et al. (2010)	BBW	PnET-BGC	Interactions between climate change
parameters and atmospheric
deposition
Phelan et al. (2016)	BBW and ForSAFE-VEG	Atmospheric deposition to northern
HBEF	hardwood forests
Tominaqa et al. (2010)
HBEF
MAGIC, PnET-BGC,
Model output comparison


SAFE, VSD

Pourmokhtarian etal. (2012)
HBEF
PnET-BGC
N cycling and climate change
BBW = Bear Brook Watershed; CL = critical load; ForSAFE-VEG = Soil Acidification in Forest Ecosystems; HBEF = Hubbard
Brook Experimental Forest; MAGIC = Model of Acidification and Groundwater in Catchments; N = nitrogen; NE = northeastern;
PnET-BGC = Photosynthesis and Evapotranspiration-Biogeochemical; SAFE = Soil Acidification in Forest Ecosystems;
SSWC = Steady-State Water Chemistry; VSD = Very Simple Dynamic.
16.1.3.2.2.1.	Acadia National Park
No modeling studies on aquatic critical loads for ACAD have been identified in the
literature.
16.1.3.2.2.2.	Hubbard Brook Experimental Forest
Tominaga et al. (2010) evaluated the performance and prediction uncertainty of four
dynamic process-based watershed acidification models: MAGIC, PnET-BGC, SAFE, and
VSD. Model output was assessed by systematically applying each of the models to data
from the HBEF. The models were used to assess future soil and stream chemistry
response to reduced levels of atmospheric S deposition. Both hindcast and forecast
predictions were qualitatively similar across the four models. Nevertheless, the temporal
patterns of projected stream ANC and soil base saturation differed. Tominaga et al.
(2010) concluded that these differences can be accommodated by employing multiple
models. Nevertheless, these results have implications for individual model applications.
Pourmokhtarian et al. (2012) used "the PnET-BGC model to evaluate the effects of
potential future changes in temperature, precipitation, solar radiation, and atmospheric
June 2018
16-30
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
CO2 on the pools, concentrations, and fluxes of major elements at the HBEF." The
climate projections were based on downscaled climate output from atmospheric-ocean
general circulation models. These climate projections indicated that the average air
temperature will likely increase at the HBEF site by 1.7 to 6.5°C over the 21st century,
with simultaneous increase in annual average precipitation ranging from 4 to 32 cm
above the long-term mean. The PnET-BGC model simulations under expected future
climate showed shifts in hydrology from later snowpack development, earlier snowmelt,
increased evapotranspiration, and a slight increase in the annual water yield. The model
results suggested that net soil N mineralization and nitrification will markedly increase
under elevated temperature. This increase resulted in simulated acidification of soil and
stream water, altering the quality of water draining from the forested watershed.
16.1.3.2.2.3.	Bear Brook Watershed
No aquatic critical loads modeling studies have been identified for BBW in the literature
since 2000.
16.1.3.2.2.4.	Other Northeastern Regions
Studies in the Northeastern region have used a variety of steady-state and dynamic
modeling approaches. The Conference of the New England Governors and Eastern
Canadian Premiers (NEG/ECP) sponsored a modeling assessment of steady-state CLs for
protection of forest soils and lakes against acidification in the Northeast region and in
eastern Canada (Ouimct et al.. 2006; Dupont et al.. 2005; Ouimet et al.. 2001). Dupont et
al. (2005) reported the SSWC steady-state aquatic CLs of acidity and associated
exceedances for lakes. Atmospheric acid loads were assessed based on atmospheric
deposition of both S and N, using a critical limit of pH 6 to protect aquatic biota. This pH
level approximately corresponds with ANC = 40 |icq/L in this region (Small and Sutton.
1986). Estimated S critical loads were exceeded in 2002 for 12.3% of all studied lakes.
Lakes having lowest calculated CLs included many in southern Vermont, eastern and
northern Maine, northern New Hampshire, and Cape Cod. Exceedances of CLs, based on
estimated acidic deposition in 2002, were highest in central and coastal Massachusetts,
southern Vermont, much of Maine, and portions of New Hampshire. Eastern Maine and
southern Vermont were notable "hot spots" where ambient S + N deposition exceeded
CLs by more than 10 meq/m2/year (Dupont et al.. 2005).
June 2018
16-31
DRAFT: Do Not Cite or Quote

-------
16.1.3.3. Integration
Table 16-10 contains a list of northeastern U.S. critical load determinations by multiple
researchers, ranging from 3-8 kg N/ha/yr (forests) to 17.5 kg N/ha/yr (herbaceous
plants).
Table 16-10 Empirical and modeled nitrogen critical loads applicable to the
northeastern U.S.
Critical Load
(kg N/ha/yr)
Ecosystem
Component
Response
N Deposition
(kg/ha/yr)
Site
Reference
5 to 12
Eastern
temperate
forest
Mycorrhizal Fungi
5.2
ACAD
Pardo et al.
(2011c)
4 to 8
Lichens
Epiphytic lichen community
change
5.2
ACAD
Pardo et al.
(2011c)
17.5
Herbaceous
Species
Increases in nitrophilic species,
declines in species-richness
5.2
ACAD
Pardo et al.
(2011c)
3 to 8
Hardwood
forest
Not specified
5.2
ACAD
Pardo et al.
(2011c)
Ellis et al. (2013)
8	Eastern	Increased surface water NO3	5.2 ACAD Pardo et al.
hardwood leaching	(2011c)
forests	Ellis et al. (2013)
ANC crucial Not specified Current legislated emissions
concentration	(= 13% reduction of SO42" by
= 20 peq/L	2015) results in limited response
Soil base	recovery
saturation	Maximum feasible technology
critical level of	reductions (= 78% reduction of
10%	SO42" by 2015) results in a more
rapid and greater extent of
chemical recovery
Not specified HBEF
Tominaqa et al.
(2010)
June 2018
16-32
DRAFT: Do Not Cite or Quote

-------
Table 16-10 (Continued): Empirical and modeled nitrogen critical loads applicable
to the northeastern U.S.
Critical Load
Ecosystem

N Deposition


(kg N/ha/yr)
Component
Response
(kg/ha/yr)
Site
Reference
CO2
Not specified
N cycling and climate change.
Not specified
HBEF (year
Pourmokhtarian
fertilization

"Under elevated temperature, net

2070-2100)
etal. (2012)
plateau of

soil N mineralization and



600 ppm

nitrification markedly increase,





resulting in acidification of soil and





stream water... Invoking CO2





fertilization effect on vegetation





under climate change





substantially mitigates watershed





N loss...Showed recovery of up to





1 pH unit (assuming pH of 6.0 as





steady-state value) and...ANC of





6.9 to 15.5 peq/L in comparison to





-3.4 peq/L mean annual





measured values (1988-2000)"



Not specified
Eastern
Interactions among climate
Not specified
BBW
SanClements et

temperate
change parameters and


al. (2010)

forest
atmospheric deposition



Not specified
Eastern
Atmospheric deposition to
Not specified
BBW and
Phelan et al.

temperate
northern hardwood forests

HBEF
(2016)

forest




Not specified
NE
Surface water NO3"
9 to 13
NE
Aber et al. (2003)

watersheds
concentrations exceeded 1 peq/L





and mean NO3" export increased





linearly with increasing deposition



<20 kg/ha/yr
Aquatic biota
Protect aquatic biota to pH 6.0
Not specified
NE
DuDont et al.
S + N (= 9.6 kg




(2005)
S042"/ha/yr)





ACAD = Acadia National Park; ANC = acid neutralizing capacity; BBW = Bear Brook Watershed; C02 = carbon dioxide;
ha = hectare; HBEF = Hubbard Brook Experimental Forest; kg = kilogram; L = liter; |jeq = microequivalents; N = nitrogen;
NE = northeastern; N03" = nitrate; ppm = parts per million; S = sulfur; S042" = sulfate; yr = year.
16.1.4.	Long-Term Ecological Monitoring
1	This section summarizes research on the long-term effects of S and N deposition in the
2	case study areas. The focus is primarily on research published since about the Year 2000.
3	The acidification and nutrient enrichment subsections are each organized by three case
4	study areas followed by relevant information for the overall Northeast region. Key
5	publications are summarized in Table 16-11.
June 2018	16-33	DRAFT: Do Not Cite or Quote

-------
Table 16-11 Summary table of observed terrestrial and aquatic acidification
long-term trends in Hubbard Brook Experimental Forest and Bear
Brook Watershed.
Variable
Species
Response
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Site
Reference
C:N
NA
Significant increase over
25 yr
Not
specified
Not
specified
HBEF
Campbell et al.
(2007)
Biomass
Sugar
maple
Decrease
Not
specified
Not
specified
HBEF
Campbell et al.
(2007)
Al+
NA
Decrease
Not
specified
Not
specified
HBEF
Campbell et al.
(2007)
Base cation
Ca and Mg
depletion
NA
Historical forest cutting had
little impact on
exchangeable cation soil
pools
Not
specified
Not
specified
HBEF
Gbondo-
Tuabawa and
Driscoll (2003)
SO42"	NA	Watershed 6, had a 32% Not	Not	HBEF
decline in the annual	specified specified 1963-1994
volume-weighted
concentration
(-1.1 [jeq/L/yr)
SCM2" + NOs" NA
Declines in stream
Not
Not
HBEF
Driscoll et al.

concentrations of strong
specified
specified
1963-1994
(2001b)

acids (-1.9 [jeq/L/yr)




Sum of base NA
-1.6 [jeq/L/yr
Not
Not
HBEF
Driscoll et al.
cations

specified
specified
1963-1994
(2001b)
pH	NA	Small but significant	Not	Not	HBEF	Driscoll et al.
increases in stream pH, specified specified 1963-1994 (2001b)
from 4.8 to 5.0
Driscoll et al.
(2001b)
Likens et al.
(2001)
Fish
community
Slimy
sculpin
(Cottus
cognatus),
blacknose
dace
(Rhinichthys
atratulus),
and brook
trout
(Salvelinus
fontinalis)
Shift in the upper mainstem
of Hubbard Brook from the
presence of at least three
fish species (slimy sculpin
[Cottus cognatus],
blacknose dace
[Rhinichthys atratulus], and
brook trout [Salvelinus
fontinalis]) to only brook
trout
Not	Not	HBEF	Warren et al.
specified specified (1960s-2007) (2008)
June 2018
16-34
DRAFT: Do Not Cite or Quote

-------
Table 16-11 (Continued): Summary table of observed terrestrial and aquatic
acidification long term trends in Hubbard Brook
Experimental Forest and Bear Brook Watershed.



Deposition
Addition


Variable
Species
Response
(kg/ha/yr)
(kg/ha/yr)
Site
Reference
Base cation
Sugar
American beech and red
8.4 N
WB
BBW
Elvir et al.

maple,
spruce had lower foliar Ca,
(wet + dry)
25.2 kg

(2006)

American
Mg, Zn concentrations;

N/ha/yr



beech, red
nutrient imbalance may

(NH4)2S04



spruce
offset potential

WB




photosynthesis benefits

28.8 S




Sugar maple higher






photosynthesis rates/no






decrease in Ca, Mg, Zn






concentrations




Base cation
Sugar
Little evidence of BC
18.5 (1980)
WB
BBW
SanClements

maple,
depletion but confounded
4.74 (2010)
25.2 kg

et al. (2010)

American
by ice storm litter
Wet S042"
(NH4)2S04



beech, red
mineralization
2.8
WB



spruce

Inorganic N
28.8 S


Mineralization
Sugar
Storm caused increased
Not
WB
BBW
Fernandez et

maple,
litterfall and accelerated
specified
25.2 kg

al. (2003)

American
mineralization, obstructing

N/ha/yr



beech, red
temporal trends in soil

(NH4)2S04



spruce
chemistry (17 yr)

WB






28.8 S


pH and ANC
NA
Decrease
Not
WB
BBW
Norton et al.



specified
25.2 kg

(2004)




N/ha/yr






(NH4)2S04






WB






28.8 S


Stream
NA
Decrease
Not
WB
BBW
Norton et al.
dissolved


specified
25.2 kg

(2004)
S042"



N/ha/yr

Fatemi et al.




(NH4)2S04

(2012)




WB






28.8 S


Base cation
NA
Increased BC export.
Not
WB
BBW
Norton et al.
export in

Export rates declined after
specified
25.2 kg

(2004)
runoff

7 yr treatment.

N/ha/yr

Fatemi et al.




(NH4)2S04

(2012)




WB






28.8 S


Al+
NA
Increase
Not
WB
BBW
Norton et al.



specified
25.2 kg

(2004)




N/ha/yr

Fatemi et al.




(NH4)2S04

(2012)
WB
28.8 S
June 2018
16-35
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
Table 16-11 (Continued): Summary table of observed terrestrial and aquatic
acidification long term trends in Hubbard Brook
Experimental Forest and Bear Brook Watershed.
Variable
Species
Response
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Site
Reference
Release of S
from internal
storage pools
to drainage
water
NA
Increased
Not
specified
0
General
(1972-2008)
Mitchell and
Likens (2011)
S042"
NA
Biogeochemical control of
SO42" export from forested
watersheds in HBEF has
shifted from atmospheric S
deposition to climatic
factors that regulate soil
moisture
Not
specified
0
General
Mitchell and
Likens (2011)
Critical loads
Varied
Varied
Not
specified
Not
specified
NE
Pardo et al.
(2011c)
ANC
NA
Regional surface water
ANC did not change
significantly in New
England during the 1990s
Not
specified
0
NE
U.S. EPA
(2003)
Al+ = aluminum; ANC = acid neutralizing capacity; BC = base cation; BBW = Bear Brook Watershed; C = carbon; Ca = calcium;
ha = hectare; HBEF = Hubbard Brook Experimental Forest; kg = kilogram; L = liter; |jeq = microequivalents; Mg = magnesium;
N = nitrogen; NA = not applicable; NE = northeastern; (NH4)2S04 = ammonium sulfate; N03" = nitrate; S = sulfur; S042" = sulfate;
W6 = Watershed 6; WB = West Bear Brook; yr = year; Zn = zinc.
Acadia National Park is not a federally funded long-term monitoring study area.
16.1.4.1. Long-Term Monitoring of Acidification
Acidic deposition has been shown to be an important factor causing decreases throughout
much of the Northeast region in concentrations of exchangeable base cations in soils,
which were naturally low historically. Base saturation values less than 12% predominate
in the B-horizon in portions of the region where soil and surface water acidification from
acidic deposition have been most pronounced (Sullivan et al.. 2006a; Bailey et al.. 2004;
David and Lawrence. 1996).
Mitchell et al. (2011) evaluated S mass balances for 15 watersheds in the northeastern
U.S. and southeastern Canada, including watersheds at HBEF, BBW, and Cone Pond in
New Hampshire, for the period 1985 to 2002. Most study watersheds showed evidence of
internal watershed sources of SO42 . likely from mineralization of organic S stored from
decades of high S deposition. Mobilization of the mineralized S contributed an estimated
June 2018
16-36
DRAFT: Do Not Cite or Quote

-------
1-6 kg S/ha/yr to the stream fluxes of SO42 . This mobilization has affected the observed
rates of recovery from acidification as S deposition has declined.
Surface water chemistry in the northeastern U.S. has been characterized, surveyed,
resurveyed, and monitored in many studies conducted over the last several decades
(Table 16-12V
Table 16-12 Example surface water acidification chemistry studies in the
Northeast case study region.




Time


Focus
Results
Deposition
Addition
Period
Site
Reference
N export
An unburned study
Not
Not applicable
Pre- vs.
ACAD
Nelson et al.

watershed exported 10 to
specified

post-1947

(2007)

20 times more inorganic N


fire



than the burned watershed;






in addition, retention of






inorganic N was 96% in the






burned watershed vs. 72%






in the unburned watershed.





N export
Total N + P exported by
Not
Not applicable
Not
ACAD
Nelson et al.

watersheds entirely within
specified

specified

(2007)

ACAD were significantly






lower than exports in






watersheds that were






partly or completely outside






the park.





Reference
Decline in stream SO42"
Not
Not specified
1963-1994
HBEF
Driscoll et al.
stream
(-1.1 [jeq/L/yr)
specified



(2001b)
chemistry
Decline in SO42" + NO3" in






stream (-1.9 peq/L/yr)





S budget
Biogeochemical control of
Not
Not specified
1965-2008
HBEF
Mitchell and

SO42" export has switched
specified



Likens (2011)

from atmospheric






deposition to climate






factors that regulate soil






moisture.





Stream
Stream dissolved SO42",
Not
1,800 eq/ha/yr
1987-2000
BBW
Norton et al.
trends for
pH, and ANC decrease
specified
(NH4)2S04


(2004)
reference
Base cation and Al+





and
increase





treatment






catchments






June 2018
16-37
DRAFT: Do Not Cite or Quote

-------
Table 16-12 (Continued): Example surface water acidification chemistry studies in
the northeast case study region.
Time
Focus	Results	Deposition Addition Period	Site	Reference
Low-flow and Base cation, NH4+and CI" Not	1,800 eq/ha/yr 1987-2006 BBW	Navratil et al.
high-flow unchanged	specified (NH4)2S04	(2010)
stream
chemistry of
reference
and
treatment
catchments
Chronic
stream
chemistry for
reference
and
treatment
catchments
Increased BC export;
declined after 7 yr of
treatment.
Decrease pH and alkalinity
Increased dissolved Al+,
NOs", S042"
Not	1,800 eq/ha/yr 1987-2007 BBW	Norton etal.
specified (NH4)2S04	(2010)
Episodic
stream
chemistry for
reference
and
treatment
catchments
"18 yr of N and S addition
have not affected the
natural drivers of episodic
acidification. The
contribution of SO42" to the
ANC decline in WB has
been increasing linearly
since the beginning of
watershed treatment while
the role of NO3" has
remained relatively
constant after an initial
increase."
Not	1,800 eq/ha/yr 1988-2006
specified (NH4)2S04
BBW
Laudon and
Norton (2010)
Lake	Ubiquitous decreases in Not	Not specified 1986 and NE region Warbv et al.
resurvey the concentration of	specified	1998	(2009)
inorganic Al across the
region.
Organic monomeric Al also
declined region-wide in
New England.
In 2001, only seven study
lakes (representing
130 lakes in the Northeast
region) that had AN above
the toxic threshold of 2 |jM,
compared with 20 sampled
lakes (representing
449 lakes in the
population) in 1986.
Chronic and	Rates of change for	Not	Not specified 1990-2000 NE and	U.S. EPA
episodic lake individual water bodies	specified	Appalachian (2003)
and stream	ranged from about-1.5 to	Mtns.
chemistry	-3 peq/L/yr.
June 2018	16-38	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
Table 16-12 (Continued): Example surface water acidification chemistry studies in
the northeast case study region.
Focus
Results
Deposition Addition
Time
Period
Site
Reference
Regional NO3" concentrations in
lake trends New England and the
Adirondack Mountains,
which had no trend prior to
2000, declined at a rate of
-0.05 [jeq/L/yr
Not	Not specified 2000-2010 New England Strock et al.
specified	and	(2014)
Adirondacks
Response to Oa-horizon BS decreased Not
deposition from 56.2 to 33.0% and specified
reduction almost equivalent changes
in carbon-normalized
exchangeable Ca and
exchangeable Al
Not specified
1984 to
2001
NE region
Warbv et al.
(2009)
Suggested a nascent	Not	Not specified 1992-1993 General Lawrence et
recovery of soil acid-base specified	vs.	al. (2012)
chemistry at some	2003-2004
locations in the Northeast
where six red spruce
stands sampled.
Exchangeable Al in the
Oa-horizons decreased
(p < 0.05) by 20 to 40% at
all New England sites.
Evidence of base cation
depletion and decreased
Ca:AI ratio
ACAD = Acadia National Park; Al = aluminum; ANC = acid neutralizing capacity; BBW = Bear Brook Watershed; BC = base cation;
BS = base saturation; Ca = calcium; CI" = chloride; ha = hectare; HBEF = Hubbard Brook Experimental Forest; kg = kilogram;
L = liter; |jeq = microequivalents; |jM = micromole; N = nitrogen; NE = northeastern; NH4+ = ammonium; (NH4)2S04 = ammonium
sulfate; N03" = nitrate; Oa = well decomposed humus; P = phosphorus; S = sulfur; S042" = sulfate; WB = West Bear Brook;
yr = year.
These studies have focused on both chronic and episodic responses of reference and (in
the case of BBW and HBEF) experimentally treated catchments.
The U.S. EPA's Long-Term Monitoring (LTM) Program has been collecting monitoring
data since the early 1980s for many lakes and streams in acid-sensitive areas of the U.S.,
including within the Northeast region. These data allow evaluation of trends and
variability in key components of lake and stream water chemistry prior to, during, and
subsequent to 1990 CAAA Title IV implementation. Throughout the Northeast region,
the concentration of SO42 in surface waters has decreased substantially, often by
one-third to one-half, subsequent to the 1990 CAAA. These declines in lake and stream
S042 concentrations were considered consistent with observed declines in S wet
deposition and have been corroborated by other studies showing that SO42
concentrations in northeastern lakes have decreased steadily since about the late 1970s
June 2018	16-39	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
[e.g., (U.S. EPA. 2003; Driscoll et al.. 1995)1. Regional surface water ANC did not
change significantly in New England during the 1990s (U.S. EPA. 2003).
The impact of the 1990 CAAA relevant to aluminum releases to surface waters was
assessed by Warbv et al. (2008). who surveyed 113 lakes in the Northeast that had been
sampled in 1986 and again in 2001. They "found ubiquitous decreases in the
concentration of inorganic Al across the region." In 2001, there were only 7 study lakes
(representing 130 lakes in the Northeast region) that had inorganic Al above the toxic
threshold of 2 (j,M, compared with 20 sampled lakes (representing 449 lakes in the
population) in 1986.
Strock et al. (2014) analyzed recent trends in wet deposition and lake chemistry using
long-term monitoring data for lakes in New England and the Adirondack Mountains.
From 2000 to 2010, lake NOs concentrations decreased at a rate of-0.05 (j,eq/L/year (no
trend before 2000). There was a shift to nontoxic organic Al. Both ANC and pH
exhibited variable trends.
Table 16-13 summarizes the effects of N deposition on watershed nutrient in the three
case studies areas and relevant Northeastern regional studies.
Table 16-13 Summary of observed terrestrial and aquatic nutrient enrichment
long-term responses in Acadia National Park, Hubbard Brook
Experimental Forest, Bear Brook Watershed, and other northeastern
regions.
16.1.4.2. Long-Term Monitoring of Nitrogen Enrichment
Variable
Species
Response
Deposition Addition
(kg/ha/yr) (kg/ha/yr) Site Reference
N export
Not specified An unburned study	Wet
Not	ACAD	Nelson et
specified (pre- vs. al. (2007)
post-1947
watershed exported 10 to	deposition
20 times more inorganic	only (value
N than the burned	not specified)
watershed.
fire)
June 2018
16-40
DRAFT: Do Not Cite or Quote

-------
Table 16-13 (Continued): Summary of observed terrestrial and aquatic nutrient
enrichment long-term responses in Acadia National
Park, Hubbard Brook Experimental Forest, Bear Brook
Watershed, and other northeastern regions.
Variable
Species
Response
Deposition
(kg/ha/yr)
Addition
(kg/ha/yr)
Site
Reference
N mass balance Sugar maple,
American
Beech,
yellow birch
Stream water export
decrease from 4 to
1 kg N/ha/yr
Shift to a net N sink
storing ~8 kg N/ha/yr;
unclear whether N is
accumulating or being
lost through
denitrification.
~7 kg N/ha/yr 0
HBEF
Yanai et
al. (2013)
DIN loss	Sugar maple, In W6, "losses were	"0.13g/m2/yr 0	HBEF	Aber et al.
American elevated in 1960s by a is 20% of	(2002)
Beech,	combination of recovery wet + dry
yellow birch from extreme drought and deposition"
a significant defoliation
event. N deposition
alone, in the absence of
other disturbances, would
have increased DIN
losses by 0.35 g N/m2/yr."
Biogeochemical Sugar maple, Modeled soil freeze-thaw Not specified 0	HBEF	Campbell
process	American to Year 2100. "Shortened	et al.
Beech,	frost covered period has	(2010)
yellow birch biogeochemical process
implications."
Biogeochemical Sugar maple, "A relatively mild freezing Not specified 0	HBEF	Groffman
process	American event induced significant	et al.
Beech,	increases in N	(2011)
yellow birch mineralization and
nitrification rates, solute
leaching, and soil N2O
production and caused
significant decreases in
soil methane uptake. Soil
freezing events may be
major regulators of soil
biogeochemical
processes and solute
delivery to streams in
forested watersheds."
NO3" export in Sugar maple, NO3" retention greater ~6 kg N/ha/yr	HBEF	Judd et al.
2006	American than expected. "Changes (1999-2008)	(2011)
Beech,	over last five decades
yellow birch have reduced impacts of
frost events on watershed
NO3" export."
June 2018
16-41
DRAFT: Do Not Cite or Quote

-------
Table 16-13 (Continued): Summary of observed terrestrial and aquatic nutrient
enrichment long-term responses in Acadia National
Park, Hubbard Brook Experimental Forest, Bear Brook
Watershed, and other northeastern regions.
Deposition Addition
Variable	Species	Response	(kg/ha/yr) (kg/ha/yr) Site Reference
NPP and	Sugar maple, Ca nutrition promoted Not specified 0	HBEF	Battles et
photosynthetic American higher aboveground NPP	(15-yr study) al. (2014)
surface area Beech,	and increased
yellow birch photosynthetic surface
area.
Nitrification	Showed greatest	8.4 kg N/ha/yr WB 25.2 kg BBW	Jefts et al.
response to N treatments, (wet + dry) N/ha/yr	(2004)
(NH4)2S04
WB 28.8 S
(3-1,4
glucosidase
(3-1,4-N-
acetylglucos-
aminidase
Sugar maple, "Greatest leaf N + P
American contents can increase
beech, Red short-term
spruce	decomposition,
accelerate production of
more humic-like water
extractable organic
matter and thereby
potentially influence the
distribution of organic
matter within the soil C
pool."
Not affected by long-term
N enrichment in aquatic
and terrestrial habitats.
Not specified 25.2 N
BBW
since 1989
Hunt et al.
(2008)
Mineau et
al. (2014)
WB vs. EB Not specified Little difference found
streams: Leaf
breakdown
Not specified Not	BBW	Simon et
specified	al. (2010)
WB vs. EB Not specified "Virtually identical"
streams:
Invertebrate
production
Not specified Not	BBW	Simon et
specified	al. (2010)
WB vs. EB NA	N uptake responsive
streams:
Nutrient uptake
Not specified Not	BBW	Simon et
specified	al. (2010)
Base cation
Sugar maple, American beech and red 8.4 kg N/ha/yr WB 25.2 kg BBW
American spruce lower foliar Ca,
beech, red Mg, Zn concentrations;
spruce	nutrient imbalance may
offset potential
photosynthesis benefits.
Sugar maple higher
photosynthesis rates; no
decrease in Ca, Mg, Zn
concentrations.
(wet + dry)
N/ha/yr
(NH4)2S04
WB 28.8 S
Elvir et al.
(2006)
June 2018
16-42
DRAFT: Do Not Cite or Quote

-------
Table 16-13 (Continued): Summary of observed terrestrial and aquatic nutrient
enrichment long-term responses in Acadia National
Park, Hubbard Brook Experimental Forest, Bear Brook
Watershed, and other northeastern regions.
Variable	Species	Response
Watershed N Varied	Varies widely. Not directly
retention	related to N loading.
Deposition Addition
(kg/ha/yr) (kg/ha/yr) Site Reference
Wet	0	Mid-Atlantic Campbell
deposition 1.8	and NE et al.
to 5.5 NOs"	forests	(2004b)
0.9 to
2.4 NH4+
Net nitrification Picea rubens Picea rubens density Not specified 0	NE: Ranch	Ross and
influences net nitrification.	Brook	Wemple
Watershed,	(2011)
Vermont
BAI
Sugar maple,
black cherry
Varied species with acid
deposition impact on soil
nutrient status.
Maple responds (growth)
to lime (Ca) addition.
Not specified
Not
specified
NE (7 sites)
(1937-1996)
Lonq et al.
(2009)
SOC response
Hardwood
Increased cumulative O-,
0.9 g N/m2/yr
5 to 15 g
NE: Harvard
Tonitto et

red pine
A-, and B-horizons C

N/m2/yr
Forest
al. (2014)


stocks of 211 g C/m2




ACAD = Acadia National Park; BAI = basal area increment; BBW = Bear Brook Watershed; C = carbon; Ca = calcium;
DIN = dissolved inorganic nitrogen; EB = East Bear Brook; g = gram; ha = hectare; HBEF = Hubbard Brook Experimental Forest;
kg = kilogram; m = meter; Mg = magnesium; N = nitrogen; N20 = nitrous oxide; NA = not applicable; NE = northeastern;
NH4+ = ammonium; (NH4)2S04 = ammonium sulfate; N03" = nitrate; NPP = net primary production; P = phosphorus; S = sulfur;
SOC = soil organic carbon; W6 = Watershed 6; WB = West Bear Brook; yr = year; Zn = zinc.
June 2018
16-43
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
16.1.5. Recovery
Research to measure and predict recovery in the Northeast indicates that while emission
control legislation has resulted in reduced S and N emissions from some stationary and
mobile sources, recovery of sensitive northeastern watersheds, including those located at
HBEF, BBW, and ACAD, has been limited. From 2000 to 2010, lake NOs,
concentrations in New England and the Adirondack Mountains, which had no trend prior
to 2000, declined at a rate of-0.05 (j,eq/L/year (Strock et al.. 2014). Research by Warbv
et al. (2008) in the Northeast suggested a nascent recovery of soil acid-base chemistry at
some locations where red spruce stands were sampled. In BBW, SanClements et al.
(2010) found little evidence of continued depletion or recovery of soil exchangeable base
cations. Mitchell and Likens (2011) observed that with declining inputs of atmospheric S,
the higher outputs of SO42 in drainage waters relative to precipitation inputs appear to be
driven by the S stored in the soil.
To simulate the future responses of soil and surface water chemistry to potential future
emissions controls at the reference watershed at HBEF, (Gbondo-Tugbawa and Driscoll.
2002) used the PnET-BGC (Photosynthesis and Evapotranspiration-Biogeochemical)
model. Simulation results suggested that emissions controls required by the CAAA will
likely not result in substantial future changes in critical indicators, such as soil base
saturation, soil solution Ca:Al, or stream ANC and Al concentrations. Tominaga et al.
(2010) assessed model outputs of four ecosystem models by systematically applying each
of the models to data from the HBEF. They predicted that current legislated emissions
reductions (13% reduction of SO42 by 2015) would result in limited recovery and that
maximum feasible technology reductions (78% reduction of SO42 by 2015) would result
in a more rapid and greater extent of chemical recovery.
The recovery of sensitive northeastern watersheds appears to be influenced by climate
change, which has become increasingly important to consider in understanding the
potential for recovery from decades of S and N deposition. For example, climatic change,
which increases annual stream discharge and watershed wetness, will potentially increase
S042 mobilization and hence may slow the recovery of both aquatic and terrestrial
ecosystems from acidification (Mitchell and Likens. 2011). Predictive modeling of
recovery should, therefore, consider climate change. Two modeling studies reported in
this case study are noteworthy for improving scientific understanding of recovery in the
Northeast. To test assumptions about interactions that are influenced by climate change,
Campbell et al. (2009) applied the PnET-BGC model in a northern hardwood forest
ecosystem at HBEF. Results suggested an increase in primary productivity due to a
longer growing season in the future, an increase in NO3 leaching due to enhanced net
June 2018
16-44
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
mineralization and nitrification, and a slight decrease in mineral weathering. In addition,
this warmer climate has contributed to significant declines in snow depth, snow-water
equivalents, and snow cover duration. This change has important implications for forest
ecosystem processes, such as tree phenology and growth, hydrological flow paths during
winter and spring, and soil biogeochemistry. Expected future climate change was also
simulated using ForSAFE-VEG (Soil Acidification in Forest Ecosystems model) and
found to cause an increase in soil base saturation, especially at BBW (Phclan ct al..
2016). The authors observed that climate had a lesser effect on simulated soil acid-base
chemistry at HBEF, likely due to the overwhelming influences of high S and N
deposition. They suggested that climate futures predicted by the Intergovernmental Panel
on Climate Change (IPCC) of increased temperature and precipitation will change plant
communities and N enrichment, counteracting the acidifying impacts of S and N
deposition on soil acid-base chemistry.
Finally, increasing attention is being given to developing and implementing practices to
restore forest productivity lost to impacts from acidification. Practices such as calcium
addition may increase recovery rates in northeastern forests. For example, Battles et al.
(2014) added calcium silicate to the HBEF paired watershed to determine whether Ca lost
from leaching due to acidification can be restored. Their results suggested tree biomass
recovery and reversal of forest decline through greater aboveground net primary
production and increased photosynthetic surface area. CO2 fertilization associated with
climate change is also being evaluated for its role in helping mitigate N loss in
watersheds (Pourmokhtarian et al.. 2012).
16.2. Adirondack Case Study: Adirondack Region of New York
16.2.1. Background
This case study is meant to identify the effects of nitrogen (N) and sulfur (S) in the
Adirondack region and the Adirondack state park in New York. This case study identifies
current acidification and nutrient status and empirical and modeled critical loads (CLs). It
is a supplement to the case study of acidification in the Adirondack region of New York
included in the 2008 NOx-SOx ISA [Section 3.2.2.4 of 2008 ISA (U.S. EPA. 2008a)l.
Further information about the effects in the Adirondack region can be found in
Appendix 4. Appendix 5. Appendix 7. and Appendix 8.
June 2018
16-45
DRAFT: Do Not Cite or Quote

-------
16.2.1.1. Description of Case Study Region
1	The Adirondack Mountains are in northeastern New York State and are densely forested,
2	have abundant surface waters, and have 46 peaks that extend up to 1,600 m in elevation
3	(Figure 16-9). This area includes the headlands of five major drainage basins: Lake
4	Champlain and the Hudson, Black, St. Lawrence, and Mohawk rivers. There are more
5	than 2,800 lakes and ponds, and more than 1,500 miles of rivers that are fed by an
6	estimated 30,000 miles of brooks and streams. The Adirondack Park has long been a
7	nationally important recreation area for fishing, hiking, boating, and other outdoor
8	activities.
June 2018
16-46
DRAFT: Do Not Cite or Quote

-------
Adirondack Park USGS National Land Cover
J Open Water (343,700 ac)
] Low Intensity Residential (6,100 ac)
3 High Intensity Residential (4,500 ac)
|Commerclal/lndustrial/Transportation (15,600 ac) |
] Bare Rock/Sand/Clay (500 ac)
| Quarries/Strip Mines/Gravel Pits (3,100 ac) [
| Transitional (Barren) (7,900 ac)
| Deciduous Forest (3,535,400 ac)
| Evergreen Forest (862,900 ac)
| Mixed Forest (606,700 ac)
3 Pasture/Hay (66,600 ac)
| Row Crops (45,900 ac)
| Urban/Recreational Grasses (1,800 ac)
3 Woody Wetlands (295.500 ac)
3 Emergent Herbaceous Wetlands (25,200 ac)
(TTT
A
rp-rn
0 4 8 Miles
Adirondack
Park Agency
April 2009
This map should
not be used for legal
jurisdictional determinations.
Figure 18-9 Map of Adirondack park land cover.
June 2018	16-47	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
The Adirondacks, particularly the southwestern Adirondacks, are sensitive to acidifying
deposition because they receive high precipitation, have shallow base-poor soils, and are
underlain by igneous bedrock with low weathering rates. The Adirondacks are among the
most severely acid-affected regions in North America (Driscoll et al.. 2003b; Landers et
al.. 1988) and have long been used as an indicator of the response of forest and aquatic
ecosystems to U.S. policy on atmospheric emissions of SO2 and NOx (NAPAP. 2011;
U.S. EPA. 1995a).
16.2.2. Deposition
Characteristics of nitrogen and sulfur deposition affecting the Adirondack region are
shown in Figure 16-10 and Figure 16-11. Figure 16-10A and Figure 16-11A show 3-year
average total deposition ofN and S for 2011-2013; Figure 16-10B shows the partitioning
between oxidized and reduced N; Figure 16-1 IB shows the 25-year-long time series for
wet deposition for NO3 , NH4+, SO42 and H+ obtained at the NADP/NTN (National
Atmospheric Deposition Program/National Trends Network) monitoring site at Whiteface
Mountain, NY (NY98). Surrounding areas in Vermont and New Hampshire and inserts
showing the coterminous U.S. (CONUS) are shown to place the depositional
environment in context. See Appendix 2.4. Appendix 2.5. and Appendix 2Jl for more
information on deposition in the U.S. Other maps showing the contributions of individual
species to dry and/or wet deposition are given in Appendix 2.7.
Data shown in the map Figure 16-10 and Figure 16-11A were obtained from the hybrid
modeling/data fusion product, TDEP (Total Deposition,
http: //nadp. sws .uiuc.edu/committees/tdep/tdepmaps/ and described in Annex 2,
Section 2.8). The time series of wet deposition is taken directly from data on the
NADP/NTN (Figure 16-1 IB). This was done to track changes in deposition since the
passage of the Clean Air Act Amendment (CAAA) because the Community Multiscale
Air Quality (CMAQ) model simulations used in TDEP extend back only to 2000.
Comparison of Figure 16-1 OA and Figure 16-11A indicates that the general pattern of
deposition of N and S is broadly similar; deposition estimates tend to be uniform
throughout the Adirondack Park. In addition, the area surrounding the park shows a high
degree of regional homogeneity but with higher values to the Southwest. Deposition of N
is not as high as in many areas of the central U.S. Deposition of S is considerably lower
than along the Ohio River Valley.
Figure 16-10B shows that the deposition of nitrogen is estimated to be mostly in oxidized
form throughout the entire park. In Figure 16-1 IB. wet deposition of all species shows
that downward trends in NO3 . NH4+, SO42 . and H+ are consistently found over the past
June 2018
16-48
DRAFT: Do Not Cite or Quote

-------
1	25 years, although the rate of decrease was variable, hi general, wet deposition typically
2	exceeds dry deposition of N and S in this area.
B
ha = hectare; kg = kilogram; N = nitrogen.
Figure 16-10 Total nitrogen deposition (A) and percentage of oxidized nitrogen
deposition (B) for the Adirondack case study area estimated by
the National Atmospheric Deposition Program Total Deposition
Science committee.
O Monitor NY #98
0	Monitor NY #20
0 Monitor Locations
1	I Adirondack Park Boundary
0 30 60
Kilomelef s
O Monitor NY #98
0	Monitor NY #20
Q Monitor Locations
1	1 Adirondack Park Boundary
N Deposition (kg-N/ha)
% Oxidized N Deposition
WW
June 2018
16-49
DRAFT: Do Not Cite or Quote

-------
A
B
Annual Wet Deposition and 3-Year Moving Average at
Site NY98:1990 - 2014
600
"i_ 500
>
400			
o
o
1988	1992	1996	2000	2004	2008	2012	2016
Year
•	nh4+
•	no3-
•	S04z-
•	H+ Lab
H+ = hydrogen ion; ha = hectare; kg = kilogram; mol = mole; NH/ = ammonium; N03 = nitrate; S = sulfur; S042" = sulfate; yr = year.
Figure 16-11 Total sulfur deposition (A) for the Adirondack case study area
estimated by the National Atmospheric Deposition Program Total
Deposition Science committee. Time series of wet deposition (B)
from the National Atmospheric Deposition Program/National
Trends Network Whiteface Mountain, NY
0	30 60
Kilometers
O Monitor NY #98
O Monitor NY #20
J Monitor Locations
II Adirondack Park Boundary
June 2018
16-50
DRAFT: Do Not Cite or Quote

-------
16.2.3.	Critical Loads and Other Dose-Response Relationships
16.2.3.1.	Terrestrial
1	This section presents post-2008 Adirondack area findings on the dose-response
2	relationships of N and S to terrestrial ecology, as well as the critical N and S loads for
3	maintaining ecosystem health. Additional relevant information for the Adirondack region
4	is summarized. The section presents findings on both empirical research and modeling
5	analyses.
16.2.3.1.1.	Empirical and Modeling Studies
6	Post-2000 findings from empirical and modeling studies of terrestrial dose-response
7	relationships and critical loads are summarized in this section. Table 16-14 summarizes
8	the body of empirical and modeling research identified.
Table 16-14 Terrestrial empirical and modeling research on the response of
nitrogen and sulfur deposition for the Adirondack Mountains.



Deposition/






Addition



Variable
Species
Response
(kg N/ha/yr)
Years
Site
Reference
Community
Redback
Increasing trends in
14.57 to 19.7
2009
12 upland
Beier et al. (2012)
richness and
salamanders,
snail community
kg/ha/yr as wet

hardwood

abundance,
calciphilic
richness and
NOs"; 17.44 to

forests

live biomass
species of
abundance, live
29.09 kg/ha/yr




snails
biomass of
as wet SO42",





salamanders, with
modeled (Ito et





increasing soil Ca.
al.. 2002)






1990-1999



Canopy tree
Sugar maple
Sugar maple basal
14.57 to 19.7
2009
12 upland
Beier et al. (2012)
basal area

area was positively
kg/ha/yr as wet

hardwood



correlated to forest
NO3-; 17.44 to

forests



floor and mineral soil
29.09 kg/ha/yr





(B-horizon)
as wet SO42",



modeled (Ito et
al.. 2002)
1990-1999
June 2018
16-51
DRAFT: Do Not Cite or Quote

-------
Table 16-14 (Continued): Terrestrial empirical and modeling research on the
response of nitrogen and sulfur deposition for the
Adirondack Mountains.
Deposition/
Addition
Variable Species	Response	(kg N/ha/yr) Years	Site	Reference
Tree basal
Sugar maple,
Sugar maple basal
NA
2008 Ha De-
McEathron et al.
area
black cherry,
area was positively

Ron-Dah
(2013)

American
correlated with

Wilderness


beech, red
mineral soil pH, and

Area in


maple, and
yellow birch basal

Adirondack


yellow birch
area was positively

Mountains



correlated with





mineral soil





exchangeable Ca.





Sugar maple basal





area was also





negatively correlated





with stream water





DOC.



Tree basal
Sugar maple,
Relative basal areas
NA
2004-2005 Adirondack
Paae and Mitchell
area
American
of sugar maple and

Park
(2008)

beech,
American basswood




American
were positively




basswood,
correlated with




and white
mineral soil




ash
exchangeable Ca;





American beech was





negatively correlated;





white ash was not





correlated.



Sugar maple
Sugar maple
Plots with lower soil
750 to
2009 Adirondack
Sullivan et al.
regeneration

base saturation did
1,120 eq/ha/yr
Park
(2013)


not have sugar maple
as N + S




regeneration (these
(NADP wet,




same plots also
CASTNET dry)




received higher N and





S deposition levels);





proportion of sugar





maple seedlings





dropped substantially





at base saturation





levels less than 20%.



Canopy
Sugar maple
Canopy vigor was
750 to
2009 Adirondack
Sullivan et al.
vigor and

positively correlated
1,120 eq/ha/yr
Park
(2013)
tree growth

with soil pH and
as N + S




exchangeable Ca and
(NADP wet,




Mg. Mean growth
CASTNET dry)


rates (BAI) were
positively correlated
with exchangeable
Ca and base
saturation at the
watershed level.
June 2018
16-52
DRAFT: Do Not Cite or Quote

-------
Table 16-14 (Continued): Terrestrial empirical and modeling research on the
response of nitrogen and sulfur deposition for the
Adirondack Mountains.
Deposition/
Addition
Variable Species	Response	(kg N/ha/yr) Years	Site	Reference
Tree basal All major Significant positive Gradient of 1984-2004 Adirondack Bedison and McNeil
area and trees	overall effect of N 3.5—7 kg	Park	(2009)
woody	deposition on tree N/ha/yr
biomass	growth from
increments	1984-2004, but
positive growth
effects only for red
maple, balsam fir,
and red spruce at the
species level.
NO3"	NA	PnET-BGC Model. Not specified 1999-2099 NE	Campbell et al.
leaching	Increase due to	(2009)
enhanced net
mineralization and
nitrification
Mineral NA	PnET-BGC Model. Not specified 1999-2099 NE	Campbell et al.
weathering	Slight decrease due	(2009)
to reduced simulated
soil moisture
(negative effect) and
increased
temperature (positive
effect)
BBW = Bear Brook Watershed; BC = base cation; C = carbon; Ca = calcium; ForSAFE-VEG = Soil Acidification in Forest
Ecosystems; ha = hectare; HBEF = Hubbard Brook Experimental Forest; IPCC = Intergovernmental Panel on Climate Change;
kg = kilogram; Mg = magnesium; N = nitrogen; NA = not applicable; ND = no data; NE = northeastern; (NH4)2S04 = ammonium
sulfate; N03" = nitrate; PnET-BGC = Photosynthesis and Evapotranspiration-Biogeochemical; S = sulfur; WB = West Bear Brook;
yr = year; Zn = zinc.
16.2.3.1.1.1. Critical Loads
1	Target loads of S deposition were calculated for 44 watersheds and extrapolated to
2	1,320 acid-sensitive watersheds in the Adirondacks using MAGIC and different soil
3	chemical indicator and thresholds of base saturation (5 and 10%), soil solution Bc:Al, and
4	Ca:Al [1.0 and 10.0 (Sullivan etal.. 201 la) I. In acomparison of target loads (out to
5	Years 2050 and 2100) and the 2002 deposition, only 11.6 to 13.5% of the watersheds
6	were simulated to be in exceedance of target loads to protect soil base saturation to 5%.
7	For target loads to protect soils to a base saturation of 10%, 79.7 to 87.5% of the
8	watersheds were in exceedance. For soil solution Bc:Al, 7.8 and 98.1% of watersheds
9	were exceeded by the 2002 deposition for target loads to protect soil solution ratios of 1.0
10	and 10.0, respectively. For soil solution Ca:Al, 44.1 to 58.2% of watersheds experienced
June 2018
16-53
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
exceedances of target loads associated with a soil solution ratio of 1.0, and 98.1% of
watersheds with target loads to protect Ca:Al ratios of 10.0 were exceeded. Further
investigations revealed that 58.2, 85.7, and 93.6% of watersheds could not obtain
threshold values of 10% base saturation, Bc:Al = 10, or Ca:Al = 10, respectively, even if
acidic deposition was held at zero, thereby demonstrating that these chemical indicator
threshold values were not useful for target load calculations using MAGIC in the
Adirondack Mountains.
16.2.3.2. Aquatic
This section presents research findings at Adirondack Mountains on the dose-response
relationships of N and S to aquatic ecology, as well as the critical N and S loads for
maintaining ecosystem health. Additional relevant information for the Adirondack region
is summarized. Both empirical and modeling studies of aquatic dose-response
relationships and critical loads are included in this section.
16.2.3.2.1. Empirical Studies
Studies in the Adirondack Mountains reviewed in the 2008 ISA demonstrated the effect
of acidification on fish species richness. Of the 53 fish species recorded in Adirondack
lakes, about half (26 species) were absent from lakes with pH below 6.0. Those
26 species included important recreational species plus ecologically important minnows
that serve as forage for sport fish (Baker etal.. 1990b). There is often a positive
relationship between pH and number of fish species, at least for pH values between about
5.0 and 6.5, or ANC values between about 0 and 50 to 100 |icq/L (Cosby et al.. 2006;
Sullivan et al.. 2006a; Driscoll et al.. 2003b; Bulger etal.. 1999).
As summarized in the 2008 ISA, lakes and streams having an ANC < 0 (ieq/L generally
do not support fish (Figure 16-12). The analysis shown in this figure suggests that there
could be a loss of fish species with decreases in ANC below a threshold of approximately
50 to 100 (ieq/L (Sullivan et al.. 2006a').
June 2018
16-54
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
12
10
(A
Ł
"o
Q. 8
W
« 6
tl
»¦ A
O *
O ^
JD 2
E
3 0-
-2 -
-200
Acute
wrn
Sf
ffl 
-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
water's pH declines from pH 7.0 to 4.2. A loss of about 12 species occurred in streams
that had a pH between 7 and 4.2. Regression across all 36 streams showed a loss of
4.6 species per unit pH decrease (Figure 16-13). Inorganic A1 toxicity was likely the main
cause of the loss of macroinvertebrates. The A1 concentration is strongly correlated with
surface water pH (as pH decreases, the solubility of inorganic A1 increases) and acid-base
balance as measured by BCS.
Studies in the Adirondacks have considered effects of species assemblages associated
with acid-impacted water bodies. Nierzwicki-Bauer et al. (2010) presented baseline
midsummer chemistry and community composition of phytoplankton, bacteria, rotifers,
crustaceans, macrophytes, and fish in 30 lakes of the southwestern Adirondacks
(Figure 16-14). Species richness in the lakes was correlated with acid-base chemistry at
all trophic levels except bacteria. The decrease in number of taxa per unit pH was similar
among groups and ranged from 1.75 (crustaceans) to 3.96 (macrophytes).
Percent et al. (2008) assessed bacterioplankton community diversity and structure in
18 Adirondack lakes across a range of acid-base chemistry using sequencing and
amplified ribosomal DNA restriction analysis of constructed rRNA gene libraries. Based
on principal components analysis, pH was positively correlated with bacterioplankton
community richness and diversity. The richness of several bacterial classes, including
Alphaproteobacteria, was directly correlated with pH. However, other environmental
factors, such as lake depth, hydraulic retention time, dissolved inorganic C, and nonlabile
(organically bound) monomeric Al, were also important in explaining bacterioplankton
community richness and diversity, suggesting that acidity is only one factor that controls
community composition.
Post-2008 findings from empirical studies of aquatic dose-response relationships and
critical loads are summarized in this section. Table 16-15 compiles the body of empirical
research identified.
16-56
DRAFT: Do Not Cite or Quote

-------
40
25
35
u. 30
m

-------
Bactcrioplankton
~ ~
y=2.4972x - 0.1353
~ ~ ~
R'=0.049
~
I,

	
— «i ¦ " ¦ ¦
w m ¦
¦
y=1.4259x-1.9141 —
R'-0.122
•125 4 75 525 5 75 6.25 6,75 7 25 7 75
Rotifers
y^3,S603x - 12.249
R*-0.6863
•<25 4 75 5.25 5.75 6,25 675 7.25
~ ~
Macrophytes
~
y-3.957x • 9.3734
~ ~
R-—0.2128
~ ~ ~
—— i ¦ *—

TV.—- ~ ~
~ ~
	 ^

~ ~*
K


Phytoplanklon
y=3.9744x - 9.0623
~ ~
*R'-0.49%
~
~ .


~ ~ *5^-1
~
~

42)	4 75	5 25	575	625	6,75	7 25
Crustaceans
30
25
20
15 |
10
5 1-
y-l,7529x- 3.4791
R-0.5873
4,25 4.75 5.25 5.75 6.25 6.75 7.25
Fish

y=3.7218x - 17-641
R!=0.5316
~
~
~
A r~
	—
425 4.75 5.25 5,75 625 6.75 725
Lake pH
Notes: they-axis denotes number of groups (bacteria), taxa (phytoplankton) or species (rotifers, crustaceans, macrophytes, fish).
The x-axis denotes the pH range that occurred in the 30 study lakes. In the bacterioplankton graph, two regressions were run on the
basis of genera (diamonds) and classes (squares).
Source: Nierzwicki-Bauer et ai. (2010't.
Figure 16-14 Species richness of biotic groups in 30 Adirondack study lakes
relative to midsummer epilimnetic pH during sample years.
June 2018
16-58
DRAFT: Do Not Cite or Quote

-------
Table 16-15 Aquatic empirical research on the response of nitrogen and sulfur
deposition for the Adirondacks.
Study
Time
Period
Focus
Notable Results
Changes to Acidity
Strock et al. 1980-2010 Trends in recovery for SO42" concentration decreased slightly faster in the
(2014)	lakes in the Northeast, 2000s than the 1990s—possibly the result of a decline
including 43 sites in the in S emissions that was twice as large in the 2000s
Adirondack Mountains. than it was in the previous decade. NO3" concentration
did not significantly fall in the 1990s (a decade which
had more modest N emissions reductions) but did
significantly fall in the 2000s, which saw large N
emission reductions.
Mitchell et al. 1984-2010 S deposition response in SO42" concentration has declined significantly along
(2013)	16 Adirondack long-term with total S deposition. However, there is a discrepancy
monitoring Lakes.	in the balance of deposition inputs and discharge
outputs. The authors conclude that internal S supplies
are likely contributing to high sulfur concentrations and
the slow recovery of lakes in the region, and that
internal S will become a more important factor as
emissions decline.
Waller et al. 1991-2007 The effectiveness of the
(2012)	Acid Rain Program and
Nitrogen Budget Program
in the 1990 CAAA in
reducing Adirondack lake
acidity for 42 lakes.
Sulfate concentration declined by 23.47% over the
study period, but nitrate concentration did not
significantly decline. ANC has increased over the study
period, with the number of lakes with ANC <0 dropping
by 46%. ANCg was higher than ANCcaic, suggesting
contribution to acidity from naturally occurring organic
acids that are not factored into ANCcaic.
Baron et al.	1997-2006 Empirical N loads in	Minimally disturbed Northeast lakes were estimated to
(2011b)	216 northeastern U.S.	have thresholds of 3.5-6.0 kg N/ha/yr for avoidance of
lakes including	acidic episodes.
Adirondack lakes.
Lawrence et al. 1985-2008 Focuses on	The pH in streams with lower ANC was more
(2011)	12 Adirondack streams responsive to the 50% reduction in S deposition over
rather than lakes.	the study period than pH in streams with high ANC.
Authors noted "a more muted recovery response" in
streams compared to what has been reported in lakes.
Inamdar and 2002-2004 Point Peter Brook	Results suggested that groundwater sources have
Mitchell (2008)	watershed in Western NY	greater impact on changes to sulfate flux in observed
and fluxes of sulfate	lakes than does deposition through precipitation,
following storms.
Warbv et al. 1986-2001 Speciation and
(2008)	concentration of Al in
113 northeastern U.S.
lakes (including the
Adirondacks) and the
effect of decreased acid
deposition.
The Adirondack lakes had the largest decline in Al
concentration of all lakes studied (Adirondacks:
2.01 pmol/L; all: 0.44). For reference, the decline in
SO42" deposition in the Adirondacks was about the
same as the decline in the region as a whole.
June 2018
16-59
DRAFT: Do Not Cite or Quote

-------
Table 16-15 (Continued): Aquatic empirical research on the response of nitrogen
and sulfur deposition for the Adirondacks.
Study
Time
Period
Focus
Notable Results
Changes to Biota
Sutherland et al.
(2015)
1984-2012
Brooktrout Lake
acidification and biotic
recovery.
Likely as a result of the 1990 Clean Air Act
Amendments, SO42" concentrations have dropped 56%
or 2 pequiv/L/yr since the 1980s. Plankton species
richness has also blossomed, though crustacean
species richness has had less impressive growth
(possibly due to pressures exerted by the predator-free
glassworm, Chaoborus). Brook trout (Salvelinus
fontinalis) are now successfully surviving and
reproducing after their reintroduction in 2005.
Josephson et al.
(2014)
1960-2011
Honnedaga Lake
watershed recovery.
There has been significant decline in SO42" and NO3"
since 2001, when consistent measurements in surface
water chemistry began. There were no notable trends
in zooplankton species richness as lake acidity dropped
over the observation period, but there was a significant
resurgence of adult brooktrout in the 2000s compared
to catches in the 1970s (a nearly sixfold increase). The
authors noted the relationship between acidity and
aluminum toxicity for fish. Recovery of adult trout could
be improved by lower acidity tributaries that would
make a friendlier environment for young-of-year trout.
Yu et al. (2011)
2003-2004
Focused on Hg
concentrations in biota of
44 lakes (not NOxSOx but
relevant to acidity).
Lower ANC in surface water was correlated with higher
Hg concentrations in the tissues offish, loons, and
zooplankton.
Nierzwicki-Bauer
et al. (2010)
1994-2006
Correlation between
species richness and
acid-base chemistry of
30 Adirondack lakes.
For each unit of pH (toward increased acidity), species
richness dropped from 1.4-1.8 (bacterial classes,
crustaceans) to above 3.5 (fish, rotifers, phytoplankton,
macrophytes). Authors classified species into a system
of brackets describing acid sensitivity or tolerance, with
most fish species being considered acid sensitive.
Baldiao et al.
(2009)
2003-2005
Macroinvertebrate
response to acidification
of 36 streams in the
Adirondack region.
Species richness was measured by acid
bioassessment profile (acidBAP) scores. In about half
of streams, acidBAP scores were moderately or
severely impacted (44%) from low pH—this increased
to about two-thirds of streams when including slight
impacts (69%).
Percent et al.
(2008)
2002-2002
Acidity and bacterial
communities in
18 Adirondack lakes.
Authors reported a significant decline in bacterial
species richness and species diversity with increasing
acidity. However, they did not find the overall bacterial
community composition to be significantly affected.
Notes: ANC = acid neutralizing capacity.
June 2018
16-60
DRAFT: Do Not Cite or Quote

-------
16.2.3.2.2.
Modeling Studies
Post-2008 findings from modeling studies of aquatic dose-response relationships and
critical loads are summarized in this section. Table 16-16 compiles the body of research
identified.
Table 16-16 Critical and target load and exceedance modeling studies in
Adirondack Mountains.
Reference	Model	Focus	Notable Results
Sullivan et al. (2012a) MAGIC and	TL for lakes in 2050 To achieve ANC = 50 peq/L in 2100, about
regional	and 2100	30% of lakes had simulated TL of S
extrapolation of	deposition <500 eq/ha/yr and about 600 lakes
model	were in exceedance.
Zhou et al. (2015b)	PnET-BGC	TL link to fish and The magnitude of simulated historical
zooplankton richness acidification represented by ANC loss ranged
from about 26 to 100 peq/L. The amount of
historical acidification of the lakes was related
to the total ambient deposition of
SO42" + NO3", the Ca weathering rate, and
the simulated preindustrial ANC in the
Year 1850. Adirondack lakes have lost fish
and total zooplankton species richness
beginning with the onset of acidic deposition.
Modeling results suggested that complete
recovery to preindustrial conditions may not
be possible for most acidified lake
ecosystems.
Zhou et al. (2015c)	PnET-BGC	Effects of biophysical Model simulations suggested that future
factors on the TL in decreases in SO42" deposition would be more
lakes	effective in increasing the lake water ANC
than equivalent decreases in NO3"
deposition. The difference was a factor of 4.6
during the period 2040 to 2050, but
decreased to a factor of 2 by the Year 2200.
Lakes that had longer hydrologic residence
exhibited less historical acidification and
therefore could achieve a higher ANC in
response to chemical recovery compared with
lakes that had short hydrologic residence
times.
Fakhraei et al. (2014) PnET-BGC	TMDLs for 128 acid- Model simulations suggested that an S TL
impaired lakes	equal to 79 eq S/ha/yr (representing a 60%
decrease from ambient deposition) would
lead to ANC recovery at a rate of
0.18 [jeq/L/yr through 2050, with reduced rate
of recovery thereafter.
June 2018
16-61
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
Table 16-16 (Continued): Critical and target load and exceedance modeling
studies in Adirondack Mountains.
Reference	Model	Focus	Notable Results
NAPAP (2011)	SSWC	Combined deposition To achieve ANC = 50 peq/L on average,
load of S and N to critical load of sulfur and nitrogen for lakes in
which a stream and the Adirondack Mountains is 1,620 eq/ha/yr
its watershed could
be subjected and still
have a surface water
concentration ANC of
50 peq/L on an
annual basis
Sullivan (2015)	MAGIC and Development and	To achieve ANC values of 50 and 20 peq/L in
regional extra- application of tools to	the Year 2050 and 2100, the TL to protect
polation of model document and	against acidification of surface waters was
quantify TL and their	exceeded throughout the Adirondack
exceedances	Mountains.
TL = target load; CL = critical load; ForSAFE-VEG = Soil Acidification in Forest Ecosystems; TMDL = Total Maximum Daily Load;
MAGIC = Model of Acidification and Groundwater in Catchments; N = nitrogen; S = sulfur; PnET-BGC = Photosynthesis and
Evapotranspiration-Biogeochemical; ANC = acid neutralizing capacity; SSWC = Steady-State Water Chemistry; VSD = Very Simple
Dynamic.
16.2.4. Long-Term Monitoring
This section summarizes research on the long-term effects of S and N deposition in the
Adirondacks.
16.2.4.1. Long-Term Monitoring of Acidification
Acidic deposition has been shown to be an important factor causing decreases throughout
much of the Northeast region in concentrations of exchangeable base cations in soils,
which were naturally low historically. Base saturation values less than 12% predominate
in the B-horizon in portions of the region where soil and surface water acidification from
acidic deposition have been most pronounced (Sullivan et al.. 2006a; Bailey et al.. 2004;
David and Lawrence. 1996).
At the time of the 2008 ISA, the region of the U.S. most thoroughly studied to ascertain
changes in acid-base surface water chemistry overtime was the Adirondack Mountains.
This has not changed. Several new studies are highlighted here that focused on
Adirondack surface waters, three on lakes and one on streams. The Adirondack Lakes
Survey Corporation (ALSC) has been monitoring Adirondack lakes for about 30 years.
This work has mainly focused on acid-base chemistry, but has also involved some fish
monitoring and measurement of parameters relevant to nutrient enrichment. Monitoring
June 2018	16-62	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
data have shown some chemical recovery from lake acidification, reflected in increased
pH and ANC and decreased inorganic A1 concentrations.
A mass balance study by Mitchell et al. (2011) for 15 watersheds located in the
northeastern U.S. and southeastern Canada suggested substantial sources of SO42 in
watershed soils. The internal S sources were attributed mainly to mineralization of S
stored in soils in response to decades of atmospheric S deposition. Mitchell et al. (2013)
studied 16 of the original Adirondack Long-Term Monitoring lakes that were monitored
between 1984 and 2010. Total S deposition significantly declined in all of the study lake
watersheds during the monitoring period. Correspondingly, significant decreases were
observed overtime (-2.14 |iinol/L/vcar) in lake SO42 concentrations. The authors
observed discrepancies in the calculated S mass balances that were associated with
discharge. These results suggested that internal S sources have become more important to
the watershed S budget as atmospheric S deposition has decreased. The watershed supply
of S042 was lower in those watersheds that contained lakes that had lower ANC. Limited
contributions from internal sources of SO42 to the low-ANC lakes will allow ANC
recovery. Mitchell et al. (2013) further concluded that the effects of future increases in
precipitation might become more important in regulating the amount of SO42 mobilized
from internal watershed sources.
Surface water ANC and pH recovery has been documented by Lawrence et al. (2013) and
Lawrence et al. (2011) in studies of Adirondack streams and lakes. Lawrence et al.
(2011) reported results of Adirondack Mountain stream monitoring from the early 1980s
to 2008. During that period, there was an approximate 50% reduction in atmospheric
deposition of S in the Adirondack region. On average, the pH increased by only 0.28 pH
units and ANC by 13 (j,eq/L in 12 streams that were monitored over a period of 23 years.
Larger increases in ANC and pH occurred in the monitored streams that had lower initial
ANC. In the north tributary of Buck Creek that had a relatively high DOC concentration,
the S042 concentration decreased between 1999 and 2008 at a rate of 2 |imol/L/ycar. In
the adjacent south tributary, which had a lower DOC concentration, the decrease in SO42
concentration was substantially smaller, only 0.73 |imol/L/ycar. Driscoll et al. (2016)
observed increases in ANC and pH and marked decreases in dissolved inorganic Al in 45
of 48 study lakes in the Adirondack region from 1982 to 2015, corresponding to
decreases in acidifying deposition.
Michelenaetal. (2016) reported changes in the water chemistry of 30 Adirondack lakes,
in response to reductions in acidic deposition from 1994 to 2012. The water quality of the
study lakes generally improved during the study period, but the responses were sporadic
and complex. Lake pH values increased until about 2002 and then fluctuated. Inorganic
Al concentrations generally decreased throughout the period of record. During the early
16-63
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
years of monitoring, the average pH increased dramatically from about 5.5 to 6.0. This
increase was followed by a period of reacidification for about 5 years, followed by
another period of increased pH. This 5-year cycle was then repeated to a muted extent,
even though acidic deposition continued to decline.
Chemical recovery from surface water acidification, and associated CL exceedance, in
the Adirondack Mountains has been accompanied by increasing concentrations of DOC
and organic acids, which have complicated and restricted recovery from acidification.
Lawrence et al. (2013) reported the chemistry of 42 Adirondack lakes from samples
collected between 1994 and 2011. They evaluated long-term changes in DOC and BCS
(which reflects the calculated ANC and includes an adjustment for strong organic acid
anions). Increases in DOC concentration during the study period contributed to organic
complexation of Al. This reduced the fraction of monomeric Al that was in the inorganic
(and potentially toxic) form from 57% in 1994 to 23% in 2011. Thus, the higher DOC
found during the latter sampling period increased the organic acidity of the lake water
and limited ANC recovery, but at the same time it decreased the toxicity of dissolved Al
to aquatic biota. Similarly, Strock et al. (2014) reported trends in recovery from
acidification of northeastern U.S. surface waters associated with Al chemistry. They
analyzed recent trends in lake chemistry using long-term data from lakes in the
Adirondack Mountains and New England. During the 2000s, the wet deposition of NO3
declined more than 50%. The lake NOs concentration, which showed no trend prior to
2000, declined subsequent to 2000 at a rate of-0.05 (j,eq/L/year. There was a shift to
nontoxic (organic) forms of Al. Despite the Al recovery, both the ANC and pH in the
study lakes failed to show evidence of appreciable chemical recovery; rather, the lakes
continued to exhibit variable trends in ANC and pH.
Waller et al. (2012) also evaluated the response of lake watersheds in the Adirondack
Mountains to changes in SO2 and NOx emissions and deposition, in this case using data
from the TIME monitoring program collected during the period 1991 to 2007. The
percentage of Adirondack lakes that were acidic decreased from an estimated 15.5 to
8.3%. Decreases in lake water SO42 concentrations, and to a lesser extent NO3
concentrations, generally were accompanied by increases in lake ANC. However, the
calculated ANC increased more than twice as much as the Gran titrated ANC. This
discrepancy is important for assessing lake recovery from acidification. It appeared to be
due mainly to increases in DOC (and associated organic anions) that accompanied
decreases in concentrations of the strong mineral acid anions, SO42 and NO3 . The
difference between calculated ANC and Gran titrated ANC was generally consistent with
trends of increasing DOC. Waller et al. (2012) concluded that assessments of surface
water recovery from previous acidification should consider differences in DOC
concentration and in the manner in which ANC is calculated and quantified.
16-64
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
Interpretation of long-term trends in Adirondack surface water chemistry, as summarized
above, has also been augmented by results of repeated surveys of the chemistry of lakes
initially surveyed several decades previously. Warbv et al. (2008) resurveyed 113 lakes
throughout the Northeast in 2001 that had previously been sampled by the U.S. EPA in
1986 to assess chemical recovery from lake acidification. Decreases in total Al and
inorganic Al concentrations were widespread and largest in the Adirondack Mountains
(-2.59 to -4.60 (imol/L). In 2001, only 7 study lakes (representing 130 lakes in the
population) had inorganic Al concentrations above a toxic limit of 2 (imol/L, compared
with 20 sampled lakes (representing 449 lakes in the population) in 1986. Thus, it was
estimated that in 2001 more than 300 northeastern lakes no longer had summer inorganic
Al concentrations at levels considered harmful to aquatic biota.
16.2.4.2. Recovery
Biological recovery from past acidification is a process affected by chemical, climatic,
biological, and hydrologic influences overtime. It may, under certain conditions, follow
chemical recovery of such water quality constituents as pH, ANC, and the concentrations
of SO42 , NO3 . inorganic Al, and DOC. Both chemical and biological recovery can, and
often does, lag behind changes in the levels of S and/or N emissions and deposition
because of chemical, hydrological, and biological processes and constraints. Studies at
Honnedaga Lake and Brooktrout Lake in the Adirondacks show evidence for return of
biota to levels approaching preacidification levels (Sutherland et al.. 2015; Josephson et
al.. 2014). In Brooktrout Lake, biological recovery of the food web structure has begun,
in part, due to reintroduction and re-establishment of brook trout in the lake. However,
ongoing biological recovery cannot necessarily be expected to conclude with the return of
the biological community to preacidification conditions. The reason is due to factors such
as differences in the rate of recovery of aquatic organisms, permanent loss of some
acid-sensitive species, and irreversible chemical and physical alterations to aquatic
environments during acidification (NAPAP. 2011).
Since the 2008 ISA, some paleolimnological studies have been conducted that
documented historical acidification and subsequent recovery (see Appendix 8). In one
study from the U.S., diatom shifts were linked to historical changes in pH at Brooktrout
Lake in the Adirondacks. Fragilariforma acidobiontica, a diatom that is often abundant
at pH <5.0, was present in lake sediments deposited since the 1950s, and shifts in
Mallomonas and Synura were also observed (Sutherland et al.. 2015). In this study,
phytoplankton and rotifer taxonomic richness showed substantial increases
(Figure 16-15) in association with pronounced decreases in lake SO42 . H+, and inorganic
Al concentrations. In contrast, species richness of crustaceans changed little. Sutherland
June 2018
16-65
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
et al. (2015) suggested that the absence of crustacean recovery may have been due to the
severity of the initial stress and continuing toxic pH conditions during some seasons,
and/or to a large population of predatory Chaoborus larvae present since the early 1980s
when fish were absent from the lake. In another study, Arseneau et al. (2016) concluded
that Adirondack lakes that were not previously acidified by acidic deposition will likely
not recover to predisturbance chrysophyte community structure because of the influence
of other stressors, including changes in climate.
Despite observed reductions in acidic deposition and improvement in water quality of
New York lakes, Baldigo et al. (2016) found no evidence of widespread or substantial
biological recovery of brook trout populations or broader fish communities in the
Adirondack Mountains. The study focused on 43 lakes sampled by the Adirondack Lakes
Survey Corporation during three time periods (1984-1987, 1994-2005, and 2008-2012).
Metrics reflecting fish species richness, abundance of fish species, and abundance of
brook trout did not change significantly over the 28-year period across the group of study
lakes despite a significant average ANC increase and a decrease in inorganic Al over
time. Fish species richness and catch of all fish species per net-night were positively
related to lake chemistry reflecting some limited degree of biological recovery. The
authors speculated that additional time may be needed for fish recolonization.
16-66
DRAFT: Do Not Cite or Quote

-------
160 -|
140
- 120
>•
1100
i
E
2." 80
ar
9/
| 60
Ł 40 -\
20
0
20
18
16
I"
^12
<
J10
O"
i 8
c
Ł 6
4
2
0
40
35
1 30
JC
u
= 25
.2
120
I 15
(a)
8 °
8
R2 = 0.88
(b)
(c)
| 10
o Mid-summer Epilimnetic S042' (peq I"1)
• Wet Deposition S042 (meq/m2/yr)
-¦O... ° § Q n
o o °	o
O o	o
o	° e;
® « a 8
® § 8 s
1 8 e
R2 = 0.78

~ Mid-summer Epilimnetic IMA (pM I1)
¦ Mid-summer Epilimnetic [H*] (peq I *)
R2 = 0.56
D o
D ~
I ¦
R2 = 0.55
¦ a
-s.t
~ .'«!¦ i J ¦ ' I j n
. R2 = 0.60
i '
A A
AAA
* i 4 4 R2 = 0.60
a Phytoplankton
* Total Plankton
84 86 88 90 92 94 96 98 00 02 04 06 08 10 12
Year
H' = hydrogen ion; IMA = inorganic monomeric aluminum; (jeq = microequivalent; |jM = micromolar; m = meter; S042" = sulfate;
yr = year.
Source: Sutherland et al. (2015).
Figure 16-15
(A) Midsummer sulfate concentration in the epilimnion of
Brooktrout Lake (o) and in annual wet deposition (•) at local
National Atmospheric Deposition Program/National Trends
Network Station NY52 from 1984-2012. (B) Midsummer
epilimnetic concentrations of inorganic monomeric aluminum (~)
and hydrogen ion (¦) in Brooktrout Lake from 1984-2012.
(C) Midsummer phytoplankton (A) and total plankton
(phytoplankton, rotifers, crustaceans) (A) species richness in
Brooktrout Lake from 1984-2012.
June 2018
16-67
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
16.3.
Southeastern Appalachia Case Study
16.3.1. Background
This case study focuses on the Great Smoky Mountain National Park and on several
closely located Class I areas in North Carolina (Joyce Kilmer Forest, Linville Gorge,
Shining Rock). These areas are representative of the southern Appalachian region and
also have been the focus of research on the ecological impacts of historical and current
NOx and SOx deposition. The case study summarizes literature published since 2008,
although older studies are included when relevant (e.g., conducted in Class I areas for the
purpose of investigating critical loads). The southern Appalachians are in the Eastern
Temperate Forests ecoregion, so critical loads for the broader Level I ecoregion are also
presented.
16.3.1.1. Description of Case Study Region
This case study focuses on the southern Appalachian Mountains, with special emphasis
on Great Smoky Mountains National Park (GRSM), located on the border between
eastern Tennessee and western North Carolina. The park consists of 2,100 km2 of
mountain ridges and valleys and ranges over an elevation of 267 to 2,025 m (Thornberrv-
Ehrlich. 2008). with peaks at the highest elevations in the eastern U.S. The wide range in
elevation, annual precipitation [140 cm in valleys and 215 cm on ridges and peaks,
(Thornbcrrv-Ehiiich. 2008)1. and temperature [growing season temperatures are 10-15°C
cooler on ridge tops and peaks than in valleys, (Lesser and Fridlev. 2016)1 within the
park has allowed distinct terrestrial and aquatic communities to form at different
elevations.
Geologically, the park is in the Blue Ridge physiographic province, where Cambrian
crystalline quartzite forms ridges, and Precambrian metamorphic rocks underlie the
valleys, particularly in the eastern part of the park. Carbonates and shales of the Valley
and Ridge province underlie valleys in the western part of the park where deposits of
dolomite provide substantial base cations to drainage water. The Anakeesta Formation
sits at the highest elevation of the Great Smoky Group, which forms the crest of the
mountains. The Anakeesta Formation contains iron sulfide minerals that contribute to soil
solution and surface water sulfate; it has outcropppings at Clingmans Dome, Newfound
Gap, and Chimney Tops, as well as at other high-elevation locations in the north-central
section of the park.
June 2018
16-68
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Soils of the Blue Ridge tend to be thin and shallow, overlying steep terrain (Thornbcrrv-
Ehrlich. 2008). Park soils are highly weathered and, particularly on mountain tops and
ridges, have substantial soil S adsorption capacity and limited base cation supply
(Elwood. 1991). Linville Gorge Wilderness is also within the Blue Ridge Province, with
soils that formed with low Ca, Mg, and K concentrations, in which base cation
concentrations have further declined in response to acid deposition (Elliott et al.. 2013).
There are 3,200 km of streams within the park's boundaries (Nichols and Langdon.
2007). Streams in the park are first through sixth order, and their chemistry is strongly
influenced by geology and storm events, as well as by acidic deposition. In 2006,
Tennessee listed 12 streams (67 km stream length) within GRSM as impaired under the
Clean Water Act Section 303d, because mean stream pH was below 6.0 (Fakhraei et al..
2016). Streams in GRSM that have ANC between 50 and 200 |icq/L (41%) are mostly
influenced by limestone weathering and tend to be located in the western portion of the
park (Neff et al.. 2009). Most stream ANCs in the park (54.2%) are in the 0-50 j^ieq/L
range. Streams that have ANC near or below 0 j^icq/L (2.4%) are influenced by
weathering of the Anakeesta Formation and are mainly located in the north-central
portion of the park (Neff et al.. 2009; Flum and Nodvin. 1995). Within the park,
hindcasting by PnET-BGC model simulations suggested that ANC in park streams in
1850 ranged from 28 to 107 (j,eq/L, with lower ANC at higher elevations (Zhou et al..
2015a). In the broader southern Blue Ridge province, MAGIC model simulations
hindcasted to 1860 suggested that 66 modeled streams had preindustrial
ANC > 30 (j,eq/L; in 2005, 30% of the streams had ANC below that level (Sullivan et al..
2011c). Streams in the Blue Ridge region with ANC < 20 (j,eq/L were underlain by
siliceous bedrock, including sandstone and quartzite (Sullivan et al.. 2007b).
Storm events cause episodic acidification in streams, with decreases in pH of 0.5 to
1.6 pH units in streams with ANC of <20 (j,eq/L under high-flow conditions (Lawrence et
al.. 2015b; Devton et al.. 2009). Stream chemistry responses to storm events varies by
elevation within the park. At high elevations, stream pH decreased by as much
0.5-2.0 pH units during storm events in 1990-2008 (Neff et al.. 2009; Cooketal.. 1994).
and streams above 975 m in elevation had significantly lower pH and ANC and higher
nitrate concentrations during storm events than did lower elevation streams (Neff et al..
2013). High-elevation streams (823-966 m elevation) in the Little Pigeon River
watershed had lower ANC and pH during storm events and higher sulfate, nitrate, and
organic acid anion concentrations than at low flow (Devton et al.. 2009). Streams in
watersheds influenced by the Anakeesta Formation also had significantly lower pH and
higher Al concentrations during storm flow (Neff et al.. 2013).
16-69
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
GRSM historically received some of the highest deposition loads of NOx and SOx in the
U.S. In the late 1980s, a monitor at 1,740 m elevation within the park estimated S
deposition of approximately 2,200 kg S/ha/yr, about half of which was occult deposition
(DW and SE. 1992); long-term trends in deposition from a monitor placed at a lower
elevation within the park are shown in Figure 16-18 and Figure 16-19. Much of the
historically deposited N and S remains stored within park ecosystems, with N stored in
soils, biomass, and coarse woody debris (Cai et al.. 2011b; Creed et al.. 2004). and S
adsorbed to soil (Fakhraci et al.. 2016; Cai et al.. 2011 b). The capacity of these
compartments for storing N and S are not infinite, and the capacity to retain N has been
exceeded in some areas (ecosystems have reached N saturation), as evidenced by
increased nitrate leaching and acidification of some surface waters. This suggests that
ecosystems are impaired by N deposition, and that there may be a lag time in ecosystem
recovery if deposition is reduced (Fakhraci et al.. 2016; Cai etal.. 201 la).
The range of elevation and precipitation within the park has allowed distinct terrestrial
and aquatic communities to form at different elevations, making the park one of the most
biologically diverse areas in North America. Furthermore, because the park remained
unglaciated during recent ice ages, it served as a refugium for northern plant and animal
species, as well as an evolutionary generator of new species. The All Taxa Biodiversity
Initiative is an ongoing project that seeks to document all species within the park's
boundaries, and as of March 2016, there were 19,260 species across all domains
documented within the park (https://www.dlia.org/smokies-species-tallv').
Land cover in this case study region is mostly hardwood forest; coniferous forest occurs
primarily at the higher elevations in and around GRSM (Figure 16-17). Distributed in
elevational zones from highest elevation to lowest, forest types in the park include
spruce-fir forest, beech-yellow birch forest, pine-oak forest, hemlock forest, and cove
hardwood forest. Different forest types may support endemic animal, plant, fungal, and
lichen species, as well as more broadly distributed species. There are 831 species of
lichens documented within the park (https://www.dlia.org/smokies-species-tallv); lichens
are some of the most sensitive organisms in terrestrial ecosystems to atmospheric
deposition (see Appendix 6.3.7). and work in Eastern Temperate forests (Cleavitt et al..
2015; Will-Wolf et al.. 2015) suggests that high atmospheric deposition can negatively
impact their condition, species richness, and community composition. The forests in the
park support rich understory plant communities, which are sensitive to deposition effects
such as shifts in mycorrhizal communities, competitive exclusion, and increases in
herbivory, all of which can lead to declines in herbaceous species richness (Simkin et al..
2016).
16-70
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
A number of threatened and endangered species occur within the park (Table 16-17).
GRSM is a key preserve for North American amphibian species (Nickerson et al.. 2002;
Hyde and Simons. 2001). including salamander species which have seen their North
American populations decline over the past half century (Caruso and Lips. 2013). The
park's high biodiversity has been recognized by its designation as a UNESCO World
Heritage Site and International Biosphere Reserve. The broader Tennessee-Cumberland
region contains the highest diversity of freshwater mussel and crayfish species and the
highest levels of aquatic species endemism in North America (Abell et al.. 1999). Within
the park, sampling between 1993 and 2001 found 90-168 genera of benthic
macroinvertebrates in 13 different streams (Milncr et al.. 2016). A national analysis of
federally listed species identified 49 aquatic and 4 terrestrial threatened or endangered
species impacted by anthropogenic N in the U.S. Fish and Wildlife Service Region 4,
which encompasses 10 southeastern states, including Tennessee and North Carolina
(Hernandez et al.. 2016).
Table 16-17 Species in the Southeast case study region that are listed as
threatened or endangered or as a species of concern.
Threatened and endangered species known or believed to occur in Great Smoky Mountains NP
Mammals
•	Myotis septentrionalis, northern long-eared bat—threatened
•	Myotis sodalis, Indiana bat—endangered
•	Glaucomys sabrinus coloratus, Carolina northern flying squirrel—endangered
Birds
•	Picoides borealis, red-cockaded woodpecker—endangered
Fish
•	Etheostoma sitikuense, Citico darter—endangered
•	Noturus baileyi, smoky madtom—endangered
•	Noturus flavipinnis, yellowfin madtom—threatened
Arthropods
•	Microhexura montivaga, spruce-fir moss spider—endangered
June 2018
16-71
DRAFT: Do Not Cite or Quote

-------
Table 16-17 (Continued): Species in the southeast case study region that are
listed as threatened or endangered or as species of
concern.
Threatened and endangered species known or believed to occur in Great Smoky Mountains NP
Plants
•	Geum radiatum, spreading avens—endangered
•	Spiraea virginiana, Virginia spiraea—threatened
•	Gymnoderma lineare, rock gnome lichen—endangered
Species believed to have been extirpated from the park
•	Canis lupus, gray wolf—endangered mammal
•	Canis rufus, red wolf—endangered mammal
•	Felis concolor couguar, eastern puma or cougar—endangered mammal
•	Erimonax monachus, spotfin chub—threatened fish
Federal Species of Concern found in the park
Mammals
•	Myotis leibii, eastern small-footed bat
•	Sorex palustris, water shrew
•	Sylvilagus obscurus, Appalachian cottontail
Birds
•	Ammodramus henslowii, Henslow's sparrow
•	Contopus cooperi, olive-sided flycatcher
•	Dendroica cerulea, cerulean warbler
•	Loxia curvirostra, red crossbill
•	Poecile atricapillus, black-capped chickadee
•	Sphyrapicus varius, yellow-bellied sapsucker
•	Vermivora chrysoptera, golden-winged warbler
Amphibians
•	Cryptobranchus alleganiensis, eastern hellbender
•	Desmognathus aeneus, seepage salamander
•	Eurycea junaluska, Junaluska salamander
Fish
•	Percina squamata, olive darter
•	Phoxinus tennesseensis, Tennessee dace
June 2018
16-72
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
Table 16-17 (Continued): Species in the southeast case study region that are
listed as threatened or endangered or as species of
concern.
Threatened and endangered species known or believed to occur in Great Smoky Mountains NP
Plants
•	Abies fraseri, Fraser fir
•	Calamagrostis cainii, Cain's reed-bent grass
•	Cardamine clematitis, mountain bittercress
•	Glyceria nubigena, Smoky Mountain manna grass
•	Silene ovata, Blue Ridge catchfly
NP = national park.
Source: National Park Service https://home.nps.gov/qrsm/learn/nature/te-species.htm
A number of disturbances in the park can interact with acidic deposition to affect
biodiversity. Invasive species alter the distribution of native species within the park; for
example, native brook trout (Salvelinns fontinalis) have been displaced from
approximately 75% of their historical range within park streams by rainbow trout
(Oncorhynchus mykiss) introduced to the park in the early twentieth century for
recreational fishing (Kanno et al.. 2017). Recent eradication campaigns of rainbow trout
and reintroduction of brook trout within a set of park streams showed that brook trout
quickly re-established populations (Kanno et al.. 2016). Both introduced and native
pathogens are also changing plant communities within the park. Fraser fir is endemic to
the Southeast and sensitive to acidic deposition, and 74% of remaining Fraser fir (Abies
fraseri) forests are within GSMNP boundaries (Kavlor et al.. 2017). The invasive balsam
woolly adelgid (Adelges piceae) killed mature stands of Fraser fir which historically
dominated the highest elevations of the park, (Kavlor etal.. 2017; Van Miegroet et al..
2007). The most recent survey of Fraser fir was conducted in the park in 2010 and found
the highest elevation stands regenerating despite the continuing presence of the adelgid,
suggesting some population recovery following reductions in acid deposition (Kavlor et
al.. 2017). Ongoing damage from the invasive hemlock woolly adelgid (Adelges tsugae)
may kill the mature hemlock (Tsuga spp.) stands which grow along streams in the park
and which regulate stream temperature by the dense shading they provide (Martin and
Goebel. 2012). Recent research suggests that N deposition may alter beech (Fagus
grandifolia) bark chemistry, making the trees more susceptible to mortality caused by
beech bark disease (attack by beech scale, Cryptococcus fagisuga, followed by infection
by fungal pathogens in the Neonectria genus), which was first detected in the park in
1993 (Cale et al.. 2017). In addition to disturbances caused by plant pathogens, shifting
wildfire frequency affects the distribution of plant communities within the park and the
wider region. In late November 2016, a fire burning at Chimney Tops spread rapidly
June 2018
16-73
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
north across the park and into the nearby cities of Gatlinburg and Pigeon Forge. The fire,
aided by drought conditions and strong winds, burned 17,140 acres
(https://inciweb.nwcg.gOv/incident/5112/). Wildfire suppression was the policy of the
Park from its founding in 1931 until 1996, when a program of prescribed burns was
instituted in order to regenerate forests of fire-dependent oak and pine species on dry
ridges (Schwartz ct al.. 2016).
Precipitation and temperature alter the responses of terrestrial and aquatic ecosystems to
N and S deposition. Between 1900 and 2011, annual precipitation increased 10 cm while
there was no overall trend in temperature over the same period of time at stations within
GSMNP (Lesser and Fridlev. 2016). At nearby Grandfather Mountain, NC, summer
temperature increased by more than 1.4°C between 1956 and 2011 (Soule. 2011).
Changes in temperature may alter species distributions. For example, native brook trout
inhabit about 970 km of stream length in GRSM, most of which is above 900 m elevation
(Neff et al.. 2009). Brook trout have disappeared from six high-elevation streams in the
Little Pigeon River watershed (Devton et al.. 2009; Neff et al.. 2009). Simulation
modeling by McDonnell et al. (2015) showed that acidification of streams may hinder the
movement of southern Appalachian aquatic species (including brook trout) to colder,
higher elevation habitats if stream temperatures rise.
Like GSMNP, Linville Gorge Wilderness (LGW) contains plant communities structured
by elevation. LGW comprises 43.9 km2 of mountain ridges and gorges that range over an
elevation of 426 to 1,250 m. Forest cover in the gorge is 43% coniferous tree species and
55% deciduous tree species, in plant communities from low to high elevation of acidic
cove, mesic oak-hickory, and xeric pine-oak or oak heath zones (Kantola et al.. 2016).
Soils in LGW formed from parent material with very low concentrations of Ca, Mg, and
K; and historic acidic deposition has further depleted these nutrients while decreasing soil
pH and increasing soluble aluminum (Elliott et al.. 2013). Experimental addition of
dolomitic limestone resulted in transient improvements in soil Ca:Al ratios, while
wildfire, which can release base cations from biomass into the soil, had no measurable
impact on Ca pools in soils (Elliott et al.. 2013). Between 1992 and 2011, forest structure
in LGW shifted towards trees in younger age classes, as outbreaks of beetles and
hemlock wooly adelgid, drought, and wildfire caused mortality in older and larger age
classes of trees (Hagan et al.. 2015).
16.3.1.2. Legal Authorities
GRSM is a Prevention of Significant Deterioration (PSD) Class I Area. The CAA
(42 USC 7470) authorized Class I areas to protect air quality in national parks over
June 2018
16-74
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
6,000 acres and national wilderness areas over 5,000 acres in an effort to preserve pristine
atmospheric conditions and air quality related values (AQRVs).
Wilderness areas in North Carolina located in the vicinity of GRSM that also have been
designated PSD Class I include Shining Rock, Linville Gorge, and Joyce Kilmer
Memorial Forest. All are managed by the USDA's U.S. Forest Service. Shining Rock
consists of 74 km2 of high elevation (1,450-1,550 m) hardwood forests dominated by
yellow birch and red maple on soils formed in gneissic bedrock. Linville Gorge is
43.9 km2 of acidic cove and slope forests at intermediate elevations (1,090-1,160 m)
dominated by chestnut oak and red maple on quartzite parent material. Joyce Kilmer
Memorial Forest is 68.1 km2 of low elevation (250-450 m) cove hardwoods, dominated
by tulip poplar, oaks, and eastern hemlock stands, on metasandstone parent material.
Additional information on these three wilderness areas is in (Elliott et al.. 2008). The
geography of the Class I wilderness areas described in this case study are shown in
Figure 16-16; further details on this study can be found in Appendix 16.2.3.2.
Class I areas are subject to the PSD regulations under the CAA. PSD preconstruction
permits are required for new and modified existing air pollution sources. Air regulatory
agencies are required to notify federal land managers (FLMs) of any PSD permit
applications for facilities within 100 km of a Class I area.26 The FLMs are entitled to
review and comment on PSD Class I permit applications with the authorized permitting
agency. Information on air quality monitoring by NPS within GRSM is available at the
GRSM Air Quality website (https://www.nps.gov/grsm/learn/nature/air-qualitv.htm).
26 http://webcam.srs.fs.fed.us/psd.
June 2018
16-75
DRAFT: Do Not Cite or Quote

-------
• Long-term monitored stream sites (42)
	States
I I Class 1 wilderness areas
National forests
i Simulated watersheds (30)
Cherokee j
I For^fet
Plsgah
National Forest
Lost Bottom Creek
Shutts Prong
Noiand Divide Creeks
ionai Forest
ining Roclf'
/ilderne^S
Great Smoky Mountains
National Park
Joyce raimer-Slickrock
Wilderness
Cherokee
National Fori
Nantahala
National Forest
Stimter
NaHonal Forest
Cohutta
Wilderness
Chattahoochee
National Forest
Source: Fakhraei et al. (2016).
Figure 16-16. Great Smoky Mountains National Park and nearby Class I
wilderness areas, with emphasis on water sampling locations
within GSMNP and critical loads for watersheds described in
Appendix 16.2.3.2.
16.3.1.3.	Regional Land Use and Land Cover
1	Land cover in this case study region is mostly hardwood forest; coniferous forest occurs
2	primarily at the higher elevations in and around GRSM (Figure 16-17). Although located
3	generally downwind of populous areas, GRSM has only a few human population centers
4	of any magnitude nearby. They include Knoxville, TN, 30 mi from the park (pop.
June 2018
16-76
DRAFT: Do Not Cite or Quote

-------
1
2
181,000); Asheville, NC, 30 mi (pop. 87,000); Johnson City, TN, 70 mi (pop. 64,000);
and Greenville, SC, 90 mi (pop. 61,188).
Southern Appalachian Case
Study Region: Land Cover
Case Study Locations
Native American
Reservations
Developed Land (increasing intensity)
| Barren Land
] Deciduous Forest
m Evergreen Forest
| Mixed Forest
	] Shrub/Scrub
I Grassland/Herbaceous
H Cultivated Crops
I Water/Wetlands
jtat ShiokV Sliiulitain-
'.SatifsiliV! Park
Figure 16-17 Land cover in the southern Appalachian Mountains case study
region.
16.3.2.	Deposition
3	The highest elevations of the park have received some of the highest rates of acidic
4	deposition in the U.S. (Weathers et al.. 2006; Herlihv et al.. 1993) due largely to the
5	location of upwind power plants, major agricultural regions, and the substantial amount
6	of annual precipitation.
June 2018	16-77 DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
Characteristics of nitrogen and sulfur deposition affecting the Great Smoky Mountains
Study Area are shown in Figure 16-18-Figure 16-21. Data shown in
Figure 16-18-Figure 16-23 were obtained from the hybrid modeling/data fusion product,
TDEP, http://nadp.sws.uiuc.edu/committees/tdep/tdepmaps/ and described earlier in
Appendix 2.1. However, the time series of wet deposition is taken directly from data on
the NADP/NTN website. This was done to use the long-term record of wet deposition to
illustrate trends in deposition since the passage of the Clean Air Act Amendment,
because the CMAQ dry deposition simulations involved in estimating TDEP total
deposition extend back only to 2000. Figure 16-18 shows the 25-year-long time series for
wet deposition for NO.-; . NHU+, SO42 . and H+ measured at the NADP/NTN monitoring
site Great Smoky Mountain NP (TNI 1), as well as the partitioning between oxidized and
reduced N Deposition of nitrogen, which is estimated to be mostly in oxidized form in the
study area. Although most of the area in Figure 16-18 is subject to N deposition in
oxidized form, there are areas, principally in northern Georgia, where N is deposited
mainly in reduced form. The graph of wet deposition of all chemical species shows that
downward trends in NO3 . NFU+, S042 , and H+ are consistently found over the past
25 years, although the rate of decrease has been irregular, with occasional increases.
Annual Wet Deposition and 3-Year Moving Average at
Site TN11:1990- 2014
Z
E
r
a
0 mo
• W.'
•W
•M,1-

.v**v
m tw iw
m xm au *u>
Yetr
~ Knoxvtlle. IN
O Monitor TW» 11
0 Monitor Lourticn
Appalvchia Study A/
-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
Figure 16-19 shows the 25-year-long time series for wet deposition of N and S in terms
of units commonly used to determine critical loads within GSMNP. Critical loads for
eutrophication are commonly expressed in terms of kg N/ha/yr. Critical loads for
acidification are commonly expressed in terms of kg S/ha/yr, or because both N and S
deposition can contribute to acidification, in terms of eq/ha/yr.
s.
2 «
s
8
Annual Wet Deposition and 1-Year Moving
Average at Site TN11:1990 - 2014
•	hf* kwn M4.'
•	4 fr om iO,'"
v«Vy* v
V * *
Annual Wet Deposition and 3-Year Moving
Average at Site TN11:1990 - 2014
z
I
a *
s
a.
9
o
' *
\	/X
I /	\
V *
' 'l '-fX
1M* 11M-
jwa jo u joit
t frftl U>b
JOM JQOC JOLI J016
eq = H+ equivalents; ha = hectare; kg = kilogram; N = nitrogen; NH4+ = ammonium; N03 = nitrate; S = sulfur; S042 = sulfate;
yr = year.
Notes: on the left, N and S deposition are shown in units of kg/ha/yr of N or S. On the right, N and S deposition are combined into
units of acidifying deposition, eq/ha/yr of S + N.
Source: NADP/NTN.
Figure 16-19 Trends in wet deposition of nitrogen and sulfur in Great Smoky
Mountain National Park, 1990-2014.
Figure 16-20 shows the 3-year average total deposition of N and S for 2011-2013.
Surrounding areas in southern states and inserts showing the coterminous U.S. are shown
to place the depositional environment in context. Comparison of maps indicates that the
general pattern of deposition of N and S is broadly similar, with higher values within the
study area than outside of it. Current rates of deposition of S are considerably lower than
along the Ohio River Valley, which might be expected given that the Ohio River Valley
is a major industrial region, and GRSM is a national park. Wet and dry deposition of S
were roughly equal across the study area. See Appendix 2.4. Appendix 25. and
Appendix 2J) for more information on deposition in the U.S. Other maps showing the
contributions of individual species to dry and/or wet deposition, based on TDEP are
given in Appendix 2.7.
June 2018
16-79
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
i«y V /

Ylrtt^*3
Tfro"****®
• ~
r
~ KnoiySc. TN
Q Monitor TN ¦ 11
Momtor Locator*
ApplteClM Study Am*
Gne"»"
¦

U *1

tlnjuuiij

ft 9 Mtt«|
¦¦¦¦*


*vv;
N* *
* # *
N*>


w#"»
KwilMdky
teiuwit^
~
r.
~ Knsivlitt. TN
O Monnw TN ( 11
1 Monitor Location*
Ap|>nl«chi
-------
10 km
Deposition in GRSM,
2000 {kg ha^-yr1)
Sulfur	Nitrogen
6.5-13.5
13.5-20.2
20.2-26.9
26.9-33.6
33.6-41.5
4.6-10.0
10.0—15.0
15.0-20.0
20.0-25.0
25.0-30.9
GRSM = Great Smoky Mountain National Park; ha = hectare; kg = kilogram; yr = year.
Source: Weathers et al. f2006).
Figure 16-21 Modeled sulfur and nitrogen deposition to the Great Smoky
Mountain National Park for the Year 2000.
June 2018
16-81
DRAFT: Do Not Cite or Quote

-------
Emission sectors
NH3 (livestock)
[ NH3 (fertilizer)
^ NH3 (natural)
U NOx (surface inventory)
| NOx (electric generating units)
| NOx (non-electric generating
industrial stacks)
~| NOx (aircraft)
	NOx (lightning)
NOx (soil)
Geographic footprint of Nr deposition
"I	I 	^	I	I	M
0.005 0.01 0.02 0.03 0.04 0.06 0.08 0.1
[kg N/ha/yr]
SM 10.4
Source: adapted from Figure 5 in Lee et al. (2016).
Figure 16-22. Annual-averaged monthly footprint of reactive N deposition in
Great Smoky Mountain National Park (10.4 kg N/ha/yr), and pie
chart of fractional contribution from emission sectors, as
estimated by GEOS-Chem adjoint model.
16.3.3.	Critical Loads and Other Dose-Response Relationships
1	The following sections describe critical loads determined for the Great Smoky Mountains
2	National Park, for the larger southern Appalachian region, or for the Level I ecoregion in
3	which the park resides, the Eastern Temperate Forests.
16.3.3.1. Empirical Studies
4	Thresholds of deposition are quantified for GRSM and the broader region. Sampling of
5	soil and streams at high elevations in the park in the mid 1990s found elevated
6	extractable inorganic N concentrations at high elevations (Garten. 2000). as well as
7	substantial nitrate leaching in streams from these watersheds (Van Miegroet et al.. 2001).
8	Because NOx deposition was determined by (Van Miegroet et al.. 2001) in the
June 2018
16-82
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
watersheds to be 32 kg N/ha/yr, Gilliam et al. (201 la) stated that the nitrate leaching
critical load would be <32 kg N/ha/yr. Additional water quality critical loads for the park
are based on models (see below). The comparable critical load set for leaching in the
Eastern Temperate Forests is 8 kg N/ha/yr, which is lower than the Gilliam et al. (201 la)
value. This critical load is set to protect hardwood forest soils from NOs leaching to
surface water and is based on a synthesis of data across northeastern forests (Aberetal..
2003). This quantity is in agreement with an earlier recommendation for Class I
wildernesses in the Forest Service's eastern region that N deposition should not exceed
5-8 kg N/ha/yr to protect terrestrial biota (Adams et al.. 1991). That report also
recommended an annual load for total sulfur deposition not to exceed 5-7 kg S/ha/yr to
protect terrestrial biota (Adams et al.. 1991). The critical loads set by Adams et al. (1991)
were based on observations across class I wildernesses in the Forest Service eastern
region; the deposition limits were based on the deposition loads of three Class I areas
where eutrophication and acidification were not observed.
A number of additional empirical critical loads have been set for the Eastern Temperate
Forests ecoregion based on protection of sensitive biota from the effects of nitrogen
deposition (Gilliam et al.. 201 la; Pardo etal.. 2011a). A critical load of >3 kg N/ha/yr
protects growth rates or survivorship of sensitive tree species, including red pine, yellow
birch, scarlet and chestnut oaks, quaking aspen, and basswood. A critical load of
4-8	kg N/ha/yr was set for Eastern Temperate Forests based on extrapolation from data
collected in Northwest forests to preserve sensitive lichen species and prevent lichen
community composition shifts towards more eutrophic species. A critical load of
5-10	kg N/ha/yr protects ectomychorrizal fungi associated with coniferous tree species
from community composition change. A critical load of <12 kg N/ha/yr protects
arbuscular mycorrhizal fungi associated with hardwood tree species from decreases in
abundance and community composition change. A critical load of <17.5 kg N/ha/yr
protects forest understory communities from herbaceous species loss and shifts in plant
community composition. These critical loads (3-17.5 kg N/ha/yr) represent protection for
a broad range of sensitive terrestrial species within the Eastern Temperate Forest
ecoregion (Gilliam et al.. 201 la; Pardo etal.. 2011a).
Two more recent studies document empirical critical loads for the broad Eastern
Temperate Forest ecoregion. Cleavitt et al. (2015) conducted surveys and used FIA
assessments of lichen species in the Northeast, with particular focus on the four Class I
wilderness areas in the region (Lye Brook, VT; Great Gulf, NH; Presidential Range—Dry
River, NH; and Acadia National Park, ME). In general, cumulative N and S deposition
over the course of the modeled period (2000-2013) was a more powerful explanatory
factor than annual N or S load for the lichen data. This work determined a critical load of
4.3-5.7 kg total N (oxidized + reduced N)/ha/yr based on lichen species richness, the
16-83
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
abundance of sensitive species (higher species richness of cyanolichens and fruticose
lichens below CL), and thallus condition. Cumulative S and N deposition were both
equally powerful predictors of thallus condition, but no S critical load was determined
(Cleavitt et al.. 2015). A study by Simkin et al. (2016) used FIA data to determine critical
loads for herbaceous species in a national assessment. The critical load for this analysis
was placed to prevent any loss of herbaceous species richness, and was determined to be
7.8-19.3 kg N/ha/yr for closed-canopy forest herbaceous communities in the Eastern
Temperate Forest ecoregion. The critical load determined for open-canopy (grassland,
shrubland, and woodland) herbaceous communities in this ecoregion was lower,
6.6-9.7 kg N/ha/yr (Simkin et al.. 2016). These critical load estimates may be too high
for GRSM, as Simkin et al. (2016) determined that critical loads tended to decrease under
conditions of low soil pH, high precipitation, and high temperatures, all of which affect
park ecosystems (Appendix 6.2.3.2).
16.3.3.2. Modeling Studies
Some of the critical loads modeled for GSMNP are aimed to protect terrestrial
ecosystems from N saturation and associated acidification. The Integrated Forest Study
sampled forests across North America, and several plots were located at high elevations
within GSMNP. A critical load of 2.5-9 kg N/ha/yr was determined for the park using
SMB modeling (OiaandArp. 1998). A decade later, Pardo and Duarte (2007) used SMB
and data from GRSM to determine critical loads for different forest types within the park.
Critical loads of 2.7-2.8 kg N/ha/yr to prevent terrestrial ecosystem eutrophication were
determined for low- and mid-elevation hardwood forests within the parks, while the
critical load for high-elevation spruce-fir forests was 3.4-7 kg N/ha/yr. Pardo and Duarte
(2007) also developed critical loads to protect GSMNP forests from acidification based
on pH, Al, and base cation threshold values. In low-elevation hardwood forests, the most
protective critical load was 1,080 eq S + N/ha/yr; in mid-elevation hardwood forest, the
most protective critical load was 1,650 eq S + N/ha/yr; and in high-elevation spruce-fir
forest, the most protective critical load was 2,000 eq S + N/ha/yr (Pardo and Duarte.
2007).
Critical loads of N and S deposition to protect aquatic ecosystems depend upon
underlying geology, elevation, and sensitivity of streams (i.e., preindustrial ANC) in the
southern Appalachian Mountains (Sullivan et al.. 201 lb). A recent application of the
PnET-BGC model was used to determine target loads to protect aquatic ecosystems in
12 streams within GSMNP from acidification. The target load to achieve a target ANC of
50 |icq/L by 2050 was deposition of 270-1,400 eq (S + N)/ha/yr in streams at <1,000 m
elevation and which did not drain watersheds within the Anakeesta Formation (Zhou et
June 2018
16-84
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
al.. 2015a). In an acid-sensitive, high-elevation watershed with drainage from the
Anakeesta Formation, no reduction in deposition was sufficient to reach a target of
ANC = 50 (j,eq/L, but a target load of 270 eq (S + N)/ha/yr resulted in ANC recovery to 0
by 2050 (Zhou et al.. 2015a). In a follow-up analysis, Fakhraei et al. (2016) and Fakhraei
et al. (2017a) determined total maximum daily loads of acid deposition for each of the
twelve 303(d)-listed stream watersheds at high elevations within the park. The target date
for recovery to pH of 6.0 was 2150 for these model runs, and critical loads ranged
between 240 and 960 eq/ha/yr of SOr + NOs + NH4+ deposition to eight of the twelve
watersheds. For the remaining four watersheds, no reduction in deposition was sufficient
to achieve pH of 6 by 2150; recovery in these streams is projected to take centuries
(Fakhraei et al.. 2017a). These target loads are in good agreement with critical loads set
for S only deposition based on MAGIC and Steady-State Water Chemistry (SSWC)
modeling for the broader southeastern Appalachian region (McDonnell et al.. 2014b).
The S only critical load was based on achieving a target stream ANC of 50 |icq/L. and for
high elevations in the park, the critical load was 0-250 eq S/ha/yr. Most of the
low-elevation area within the park had a critical load of 250-500 eq S/ha/yr by this
method (McDonnell et al.. 2014b).
16.3.4. Characterization and Long-Term Monitoring
The majority of studies conducted in this case study region have evaluated the effects of
acidification on ecosystems rather than the effects of eutrophication (see Table 16-18).
Acidification has been considered to be the dominant result of air pollution stress in the
region. Biogeochemical processes that govern watershed responses to acidification are
generally similar between the Northeast and the southern Appalachian Mountains in most
respects. The major exception is that soils in the southern Appalachians are unglaciated
and therefore more weathered, with greater sulfate adsorption capacity, than soils in
northeastern forests. Cai et al. (2010) used long-term monitoring data to generate an
input-output budget to identify processes that have influenced stream pH and ANC at the
Noland Divide watershed in Great Smoky Mountains NP during the period 1991-2006.
The majority (about 61%) of the net SO42 entering the watershed was retained,
confirming that S adsorption on soil is an important biogeochemical process in this
watershed. However, during large precipitation events, SO42 in wet deposition moved
more directly and rapidly to streams, contributing to episodic stream acidification.
Fakhraei et al. (2016) project that GRSM streams in watersheds with a low capacity to
adsorb SO42 and low N retention will respond positively (increase stream pH and ANC)
to additional reductions in S and N deposition, while streams in watersheds with high
S042 adsorption and low N retention (typically higher elevation watersheds) are less
June 2018
16-85
DRAFT: Do Not Cite or Quote

-------
1	responsive (or unresponsive) to reduced S and N deposition. In these watersheds,
2	increases in soil pH due to reduced deposition are projected to increase desorption of
3	accumulated SO42 from soils, resulting in a SO42 pulse to streams that will slow
4	recovery of stream ANC and pH (Fakhraci et al.. 2016). Monitoring data from the Great
5	Smoky Mountains NP indicated that the high S absorption in watershed soils delayed
6	recovery from previous stream acidification. Stream chemistry at 42 monitoring sites in
7	the park did not show substantial changes over the recent period of long-term monitoring,
8	1991-2014 (Fakhraci etal.,2016V
Table 16-18 Example soil, terrestrial biota, and surface water acidification
characterization and long-term monitoring studies in the southern
Appalachian Mountains region.
Study
(HERO ID)
Location
Time
Period
Focus
Results
Soil studies
Rice et al. (2014)
Southeastern
U.S.
2006-2021
Sulfate mass balances
for 27 forested
watersheds to estimate
cross-over years
(change from retaining
to releasing SO42").
Documented extensive S
adsorption on soils
Lawrence et al.
(2015b)
AT corridor,
including
through
GRSM
2010-2012
Soil sampling along AT
corridor and compilation
of other existing soil
chemistry data
Documented occurrence of low
soil base saturation along the
AT corridor, including within
GRSM
Cai et al. (2011a)
Noland Divide,
GRSM
2008
Laboratory soil columns
and in situ lysimeters to
determine response of
soil solution to reduced
acidic deposition.
Recovery of soil water, as
represented by renewed Ca2+
and Mg2+ retention, was
estimated to occur at S
deposition reduction of 61% of
ambient.
Cai et al. (2012)
Noland Divide,
GRSM
2009
Four sites along
elevational gradient.
Base saturation values <7% and
Ca:AI ratio <0.01 indicated high
acid sensitivity.
Elliott et al. (2008)
Three FS
Class I areas
in North
Carolina

NuCM model
Low Ca:AI (-0.3) in A-horizon at
Shining Rock and Linville Gorge
suggested stress to forests from
soil acidification.
June 2018
16-86
DRAFT: Do Not Cite or Quote

-------
Table 16-18 (Continued): Example soil, terrestrial biota, and surface water
acidification characterization and long term monitoring
studies in the southern Appalachian Mountains region.
Study

Time


(HERO ID)
Location
Period
Focus
Results
Terrestrial biota studies
Hames et al. (2002)
Assessment of
1995-1999
Wood thrush breeding
Low breeding success in the

650 study sites

success
wood thrush strongly correlated

across the


with low soil pH

range of the




wood thrush in




the eastern U.S.



McNultv and Boqas
Western NC
1999-2002
Red spruce and southern
CL for protecting forest
(2010)


pine beetle
ecosystems may not accurately




reflect risk to ecosystem health




due to multiple stresses,




including climate change, insect




infestation, and N supply.
Surface water studies
Scheffe et al. (2014)
U.S., including

Critical load by
AAI model

southern

ecoregion, as reflected


Appalachian

in AAI


Mts. region



Sullivan et al.
SAMI region

MAGIC model
Streams exhibited broad range
(2004): Sullivan et



of responses to changes in
al. (2002)



future S deposition, including




pronounced base cation




depletion. Recovery from past




acidification expected to be slow




and gradual.
Sullivan et al.
Southern
-2,000
Delimited high-interest
Lithology and elevation
(2007b)
Appalachian

area for water
explained locations of low-ANC

Mts.

acidification based on
streams.



geology and elevation,




which included almost




all known low-ANC




(<20 peq/L) streams in




the region.

Sullivan et al.
Southern

MAGIC model
Estimate target loads of S
(2011b)
Appalachian


deposition to protect stream

Mts.


ANC to multiple levels (0, 20,




50, 100 peq/L) at multiple points




in time (2020, 2040, 2100). TLs




ranged from 0 to values many




times higher than ambient S




deposition.
June 2018
16-87
DRAFT: Do Not Cite or Quote

-------
Table 16-18 (Continued): Example soil, terrestrial biota, and surface water
acidification characterization and long term monitoring
studies in the southern Appalachian Mountains region.
Study
(HERO ID)
Location
Time
Period
Focus
Results
Sullivan et al.
(2011c)
Southern Blue
Ridge
province of
southern
Appalachian
Mts. region
MAGIC model
Modeled 66 stream watersheds,
all of which were simulated to
have had preindustrial
ANC > 30 peq/L. In 2005, 30%
of streams had ANC below this
level. Median stream lost
25 peq/L of ANC between 1860
and 2005.
Hops and McMahon Southeastern
(2009)	U.S.
SPARROW model	Examined N transport relative to
landscape characteristics. N
transport coincided with Level III
ecoregion boundaries. Lower
fraction of N input delivered to
stream at locations where
primary flow path is shallow.
Cai etal. (2011b) Noland Divide, 1991-2007 Stream water chemistry Volume-weighted NO3"
GRSM	trends.	concentration decreased
0.56 [jeq/L/yr; stream pH and
SO42" concentration did not
change.
Cai et al. (2010) Noland Divide, 1991-2007 Developed input-output About 61% of incoming SO42"
GRSM	budget.	was retained; during rain events,
SO42" moved more readily to
streams.
Lawrence et al.
(2015b)
AT corridor,
including
through
GRSM
2010-2012
Sampled more than
200 streams along the
AT corridor, including
under both high- and
low-flow conditions.
Documented wide variability in
stream ANC along the AT
corridor.
Robinson et al.	Noland Divide, 1990-1999 Developed MLR model
(2004)	GRSM	of water chemistry.
Showed decrease in ANC over
time.
Robinson et al.
(2008)
GRSM
1993-2002
Conducted trends
analysis for 90 stream
sites, which showed
decreases in pH and
SO42" concentrations.
Extrapolation over 50 yr
suggested pH <6.0 for nearly all
study streams in the future.
Neff et al. (2013)
GRSM
2008-2009
Examined relationships
between stream
chemistry at base flow
and storm flow vs. basin
characteristics.
Following precipitation, pH
decreased and Al increased.
June 2018
16-8
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
Table 16-18 (Continued): Example soil, terrestrial biota, and surface water
acidification characterization and long term monitoring
studies in the southern Appalachian Mountains region.
Study	Time
(HERO ID)	Location	Period	Focus	Results
Peyton et al. (2009) Little Pigeon	2006-2007 Characterized chemistry	During storm flow, stream pH
River	of three high-elevation	and ANC decreased: SO42",
watersheds,	streams during	NO3", and organic acids
GRSM	episodes.	increased.
AAI = Aquatic Acidification Index; Al = aluminum; ANC = acid neutralizing capacity; AT = Appalachian Trail; Ca = calcium;
CL = critical load; FS = Forest Service; GRSM = Great Smoky Mountains National Park; L = liter; |jeq = microequivalents;
MAGIC = Model of Acidification of Groundwater in Catchments; Mg2+ = magnesium; MLR = Multiple Linear Regression models;
N = nitrogen; N03" = nitrate; NuCM = Nutrient Cycling Model; S = sulfur; SAMI = Southern Appalachian Mountains Initiative;
S042" = sulfate; SPARROW = Spatially Referenced Regressions on Watershed Attributes; TL = target load; yr = year.
16.4. Tampa Bay Case Study
16.4.1. Background
This case study provides an overview of the Tampa Bay estuary's ecology and research
related to nutrient impairments and nitrogen (N) deposition. It also reports on the
improvements to the bay's ecology being realized from strategies to reduce nutrient
loading to the bay. Years of historical data going back to the 1950s and unique
partnerships among government, industry, and citizen groups make Tampa Bay an
interesting example of the varied issues involved in monitoring and managing water
quality and N loading in an urban estuary. During the 1970s and early 1980s eutrophic
conditions were commonly observed in Tampa Bay. Contrary to many other estuaries
suffering from eutrophication worldwide, Tampa Bay appears to be recovering since the
mid 1980s as a result of the nutrient management strategy defined by the community
(Greening et al.. 2014; Morrison et al.. 2011). These approaches have included upgraded
sewage treatment, reduction of atmospheric N emissions, and controls on stormwater and
point sources. Atmospheric deposition continues to be a source of N to Tampa Bay and,
with the decreases in N loading from point sources, represents a greater proportion of
total N loading to the bay (Greening et al.. 2014). Estimates from both modeling and
measurement indicate that direct atmospheric N deposition to Tampa Bay is between 14
and 30% while total (direct plus indirect) atmospheric deposition is between 35 and 71%
(Poor etal.. 2013b; Poor et al.. 2013a; Greening and Janicki. 2006).
June 2018
16-89
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
Information regarding sulfur deposition is not included in this case study because N is a
primary contributor to nutrient enrichment and the focus of the monitoring studies used to
compile this case study.
16.4.1.1. Description of Case Study Region
The Tampa Bay estuary (Figure 16-23) is located on the eastern shore of the Gulf of
Mexico in Florida. At more than 1,000 km2, it is Florida's largest open water estuary,
with an average water depth of just 4 m and annual average rainfall rate of more than
125 cm (TBEP. 2015; Poor et al.. 2013a; TBEP. 2012b). It lies in the Northern Gulf of
Mexico marine ecoregion near the South Florida/Bahamian Atlantic marine ecoregion
(Wilkinson et al.. 2009). This transition zone between warm-temperate and tropical
ecoregions, combined with its shallow waters, large size, and gradient of fresh water to
salt water, allows the bay to support a diversity of organisms and habitats (Greening et
al.. 2014; USGS. 2011; USFWS. 1990. 1988).
Seagrass, a type of submerged aquatic vegetation (SAV), and mangrove forests are the
most prominent estuarine habitats within Tampa Bay (TBEP. 2015). Tidal marshes and
mud flats are also found throughout the estuary (USGS. 2011). Collectively, these
habitats support a large biodiversity of flora and fauna and play a role in nutrient cycling.
Seagrasses stabilize sediments, are a food source for wildlife, provide habitat, and serve
as a nursery for large numbers of fish and shellfish species. (Greening et al.. 2014).
Mangrove roots help to stabilize the shoreline and provide nursery habitat for important
fishery species. Mud flats and salt marshes in the bay provide habitat for wading birds,
while oyster reefs near river mouths serve to naturally filter the water and attract
recreational fish species, making them popular fishing spots (TBEP. 2015). Federally
endangered West Indian manatees (Trichechus manatus), a migratory species, visit
Tampa Bay seasonally to feed in the seagrass meadows. Other endangered species in the
bay documented to be impacted by N enrichment include green turtle (Chelonia mydas)
and Atlantic sturgeon [Acipenser oxyrinchus; (Hernandez et al.. 2016)1. Three national
wildlife refuges (Egmont Key, Passage Key, and Pinellas) are located in Tampa Bay and
support populations of nesting birds.
June 2018
16-90
DRAFT: Do Not Cite or Quote

-------
Hillsborough
OldNft .
Tampa ^
Bay |fli8-P
^Gandyl
Bridge J"
Hillsborough
Bay
Alalia
Pinellas
,g Middle
* Tampa
f Bay
lower
Tampa
Bay ¦
Manatee River
Manatee
Pasco
20
ZD Kilometers
Source: Poor et al. (2013a).
Figure 16-23 Tampa Bay mainstem segments arid watershed. Also shown are
the National Atmospheric Deposition (NADP) National Trends
Network wet deposition monitoring site FL41 at Verna wellfield in
Sarasota County and the NADP Atmospheric Integrated Research
Monitoring Network site FL18 in Hillsborough County.
1	The Tampa Bay watershed, located in the Eastern Temperate Forest ecoregion, is highly
2	urbanized. It is now the second largest metropolitan area in Florida with a population of
3	over 2 million people (TBEP. 2015). Pinellas, Hillsborough, and Manatee counties border
4	directly on Tampa Bay while Pasco, Polk, and Sarasota counties are also within the
5	estuary's 6,000-km2 watershed (Poor et al.. 2013a). Activities in Tampa Bay linked to the
6	local and regional economy include shipping, tourism, and commercial and recreational
7	fishing (Poor et al.. 2013a; TBEP. 2006; Tomasko et al.. 2005). More than 80 miles of
June 2018
16-91
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
deep-water shipping channels carry water traffic to the three seaports along the bay's
borders, in Tampa, St. Petersburg, and in northern Manatee County. The largest of these,
the Port of Tampa, consistently ranks among the busiest ports in the nation.
Mangroves, salt marshes, and SAV habitats in Tampa Bay have all experienced
significant reductions in extent since the 1950s, due to physical disturbance (dredge and
fill operations) and water quality degradation (TBEP. 2015). Excess N enrichment in the
bay is one of the main factors impacting water clarity and quality. The excess N leads to
increased algal biomass and reduced light availability for shallow-water (less than a 3-m
depth) SAV (TBEP. 2015). Numerous studies have shown that the waters of Tampa Bay
are strongly N limited, due to a high concentration of phosphorus (P) from natural
leaching and from active P mining in the watershed (Greening et al.. 2014). Therefore,
the primary focus for managing eutrophication in Tampa Bay has been the reduction of N
loading to the bay (Greening et al.. 2014).
Tampa Bay was designated an "estuary of national significance" by Congress in 1990. In
1998, the Tampa Bay National Estuary Program (TBNEP) partnership (consisting of the
Florida Department of Environmental Protection [FDEP]; the U.S. Environmental
Protection Agency; the counties of Hillsborough, Manatee, and Pinellas; the cities of
Tampa, St. Petersburg, and Clearwater; and the Southwest Florida Water Management
District) signed a formal Inter-local Agreement in support of a Comprehensive
Conservation and Management Plan (CCMP) for Tampa Bay. Upon adoption of the
Inter-local Agreement, the TBNEP became simply the Tampa Bay Estuary Program
(TBEP). TBEP continues to finance research projects in support of the CCMP and
coordinates the overall protection and restoration of the bay with assistance and support
from its many formal and informal partners (TBEP. 2015; Morrison et al.. 2011). TBEP
also coordinates the Tampa Bay N Management Consortium (TBNMC), a public-private
partnership working to reduce nitrogen loading to the bay and whose members include
local governments (Tampa, St. Petersburg, and others) along with key industries
bordering the bay, such as electric utilities, fertilizer manufacturers, and agricultural
operations.
The Tampa Bay watershed's human population is expected to increase over the next
decade and a 7% increase in N loading is projected to occur (TBEP. 2012a). To account
for these increases, the TBEP and TBNMC have adopted anew 17-ton/year reduction
target for total N loading, to offset expected increases in total nitrogen (TN) loading and
maintain TN loading rates at average annual rates for 1992-1994.
16-92
DRAFT: Do Not Cite or Quote

-------
16.4.1.2. Class I Areas
1	The Tampa Bay area is not a Clean Air Act Prevention of Significant Deterioration (PSD)
2	Class I area.
16.4.1.3.	Regional Land Use and Land Cover
3	Land use within the Tampa Bay watershed is mixed among undeveloped, agricultural,
4	urban, and mining uses, including phosphate mining and fertilizer manufacturing.
5	Figure 16-24 shows the land coverage within the bay's border communities and
6	watershed.
June 2018
16-93
DRAFT: Do Not Cite or Quote

-------
Source: Sherwood et al. (2016).
Figure 16-24 Tampa Bay overview map highlighting watershed development
and land use.
June 2018
16-94
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
16.4.2. Deposition
Both direct atmospheric deposition of N to surface waters and indirect deposition to the
watershed with subsequent transport to the Tampa Bay represent important sources of N
to the case study area. In a 25-year time series with deposition data from the nearest
National Atmospheric Deposition Program (NADP) National Trends Network (NTN)
monitoring site (Figure 16-23. Verna wellfield in Sarasota, FL), wet deposition of NO, .
NH4+, SOr . and H+ show strong interannual variability, but downward trends in wet
deposition of all these species are consistently found over the past 25 year (Figure 16-25).
Annual Wet Deposition and 3-Year Moving Average at
Site FL41:1990-2014
• /.
\ /\
s* , ' N.
• >
• 4
* *
400
« 550
L
T 300
to
— 250
o
200
c
o
*4= 150
Q. 100

-------
1	impact of air quality regulations (Poor et al.. 2013a). Although total nitrogen (TN)
2	loading to Tampa Bay has trended downward over the past 20 years, the percentage of N
3	from different sources has fluctuated. Direct atmospheric deposition has accounted for a
4	greater average percentage of the TN loading in recent years (from 2007-2011) than in
5	previous years KGreening et al.. 2014; Poor et al.. 2013a); Figure 16-261. Mainstem
6	segments of the bay vary widely in the percentage of total N loading attributable to direct
7	atmospheric deposition (Janicki Environmental. 2013).
Worst case: (~1976) 1985-1989	1990-1999
Time Period
2000-2011
Source: Greening et al. (2014).
Figure 18-26
Estimated annual loads of total nitrogen from various sources to
Tampa Bay summarized from 1976 to 2011.
BRACE also investigated the primary sources of N deposition to the bay, which included
power plants and mobile sources such as cars and trucks. Power plants had recently
June 2018
16-96
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
reduced N emissions through upgrades made as a result of TBEP's work with the
TBNMC (Poor et al.. 2013a'). The sensitivity results suggested that, per unit of emission,
mobile sources had a disproportionately higher contribution than power plant sources to
atmospheric N deposition to Tampa Bay (over the watershed, the mobile NOx emissions
were responsible for four times more oxidized-N deposition than the power plant
emissions; over the bay, the mobile NOx emissions were responsible for twice as much
oxidized-N deposition as the power plants). According to BRACE, reductions in N
emissions from mobile sources must be a part of the strategy to reduce N loading to the
bay, and that control of atmospheric N emissions both within and outside the Tampa Bay
watershed is important to restore and maintain good water quality and a healthy
ecosystem within the bay (Poor et al.. 2013a).
The U.S. EPA recently modeled atmospheric deposition in Tampa Bay using TDEP with
monitoring data from 2000 to 2013 (see Appendix 2.7). The general pattern of wet + dry
deposition of N (NOy and NHx) for 2011-2013 shows that fluxes are higher in the
northern half of the study area than the southern half (Figure 16-27A). The deposition of
N varies by roughly a factor of two across the study area with higher deposition of N to
the north and west of Tampa Bay and a maximum in the northeast corner of the study
area. Fluxes are typically higher than those for the rest of Florida. However, they are not
as high as in many areas of the central and northeastern U.S. Dry deposition of NO2
accounts for approximately 1/6 of total deposition of oxidized N. At Tampa Bay (and
some other NADP sites), dry fluxes are directed upward using the Community Multiscale
Air Quality (CMAQ) model, due to the implementation of bidirectional exchange of NH3.
(However, the net flux of NHx in CMAQ is still directed downward from wet
deposition.).
Most N deposition in the case study area except for the northeastern corner is estimated
to be mostly in oxidized form throughout the study area and in most of Florida, which is
consistent with many areas in the eastern and southern U.S. (Figure 16-27B). This is
generally in agreement with findings from BRACE modeling where oxidized forms of N
made up 60% of the total atmospheric loading to Tampa Bay, compared to 40% for
reduced forms such as ammonia [NH3; (Poor et al.. 2013a; TBEP. 2012bVI. In the study
area dry deposition dominates over wet deposition of N, which contrasts with many areas
in southern Florida where wet deposition dominates (Poor et al.. 2013a).
16-97
DRAFT: Do Not Cite or Quote

-------
A. Total Wet and Dry N Deposition
in Southern Florida 2011-2013
(12 km resolution)
B. Percent Oxidized N Deposition in
Southern Florida 2011-2013
(12 km resolution)
0	Monitor FL# 41
Monitor Locations
1	I Tampa Bay Study Area
ha = hectare; kg = kilogram; km = kilometer; N = nitrogen.
Notes: the Verna Wellfield National Atmospheric Deposition Program/National Trends Network monitoring site, FL41, was chosen to
characterize long-term wet deposition of N species, because it is closest to the Tampa Bay study area. The site is located at the
southern edge of the study area in Sarasota, FL and is shown as the grey dot on the maps.
Source: data shown in the figures were obtained from the hybrid modeling/data fusion product, total deposition,
http://nadp.sws.uiuc.edu/committees/tdep/tdepmaps/. and described earlier in Appendix 2.7.
Figure 16-27 (A) Wet and dry nitrogen deposition in Tampa Bay and the
surrounding area. (B) Percentage of oxidized nitrogen deposition
in Tampa Bay and the surrounding area.
16.4.3.	Long-Term Ecological Monitoring
1	Tampa Bay has been intensively studied over the years due to its significance as a natural
2	resource and its susceptibility to degradation from eutrophication. There is monthly
3	ambient water quality data going back to 1972 (Sherwood et al.. 2016). Phytoplankton
4	biomass and primary production have been assessed regularly in the bay from the 1980s
June 2018	16-98	DRAFT: Do Not Cite or Quote
N Deposition (kg-N/ha)
0	Monitor FL#41
0 Monitor Locations
1	I Tampa Bay Study Area

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
and benthic organisms since 1993 (Greening et al.. 2014). Aerial photography of the
Tampa Bay shoreline is available from the early 1950s showing the extent of seagrass
coverage prior to rapid human population increases in the watershed (Greening et al..
2014). A consistent monitoring program of seagrass extent (every 2 year) was started in
1988 (Greening et al.. 2014).
Eutrophication symptoms were observed in Tampa Bay as early as the 1970s. These
included phytoplankton and macroalgal blooms and areas of low dissolved oxygen
(Greening et al.. 2014). According to Bricker et al. (2007). algal blooms were common in
the 1970s and caused foul odors and aesthetic impairment. In the late 1970s, hypoxic and
anoxic conditions along the shoreline of the most urbanized segment of Tampa Bay
caused extensive mortality of benthic organisms in the late summer months (Greening et
al.. 2014). Fish kills have also been observed in and near Tampa Bay. Historically, the
most obvious symptom of eutrophication in Tampa Bay was the loss of light at depth due
to elevated algal biomass. The loss of SAV from reduced light availability was dramatic.
In 1950, approximately 16,000 hectares of SAV were present. By the early 1980s, over
half of this area had been lost rFigure 16-28 (Bricker et al.. 2007; Haddad. 1989)1.
Tampa Bay is somewhat unique in that the bay waters have historically exhibited a high
level of P from local phosphate deposits, P mining, and fertilizer manufacturing and
shipping operations. The high phosphorus concentrations have resulted in an unusually
low N:P ratio in comparison to other estuaries, making N the limiting nutrient for
phytoplankton growth (Greening et al.. 2014; Greening and Janicki. 2006). Numerous
nutrient limitation studies using natural Tampa Bay phytoplankton populations and
cultured test organisms have confirmed N limitation (Greening et al.. 2014).
16-99
DRAFT: Do Not Cite or Quote

-------
f

t c—S
j ^ 1 rl
\V ^ # n
% ,
\ .\
*3r
4
Mi vv
¦
Ł
V. * TTpa
\2 a, Bay f>
\J0y J)
u Jt{
*?c
/
( ^ ' ' VNyv/
v f 1 ^
. *»J-V.'
*>t N 	 i
H
¦4*
10 km

25
iim
km = kilometer; mi = mile.
Notes: red indicates area of seagrass iost from 1950 to 1990.
Source: Bricker et al. (2007).
Figure 16-28 Seagrass Loss in Tampa Bay.
16.4.3,1. Indicators of Enrichment and Eutrophication in Tampa Bay
Indicators of historical water quality trends in the bay include chlorophyll a concentration
and seagrass coverage (Greening et al.. 2014).
June 2018
16-100
DRAFT: Do Not Cite or Quote

-------
16.4.3.1.1. Chlorophyll a
1	Phytoplankton biomass as measured by chlorophyll a concentration is one of the most
2	effective indicators of eutrophication in the bay (Sherwood et al.. 2016). Chlorophyll a is
3	directly linked to nutrient inputs and has been measured in the bay since 1972. Increased
4	algal biomass results in reduced light penetration in the water, thereby harming seagrass
5	and ultimately resulting in the loss of SAV area.
6	The TBEP has developed water quality models to quantify linkages between N loads and
7	bay water quality. N is generally the primary limiting nutrient, and chlorophyll a
8	variation in the bay responds most significantly to watershed TN loads ITigure 16-29
9	(Greening and Janicki. 2006)1.
Tampa Bay Estuary Program
Nitrogen Loading - Chlorophyll a Relationship
Comparison of Predicted and Observed Chlorophyll a Concentrations

Predicted



40


86-98
99-07
REF
30

+

20

4*4 *o yf -f o ~
D o i>p9, ~ Q,e b
*¦ a 00 % ' * P ~ ~ 10
~ % at J* - t , . +*
\ * ° o

10
• >
' f*n> '3 •
ikdwEBKafip**- >*o 0
iB& v-1 *

0



3
10 20 30
Observed
40
Source: Janicki Environmental (2011).
Figure 16-29 Comparison of observed chlorophyll a and that predicted from
the total nitrogen load—chlorophyll a relationships for all four
mainstem Tampa Bay segments, for 1986-1998 and 1999-2007.
June 2018
16-101
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
Data show that, with the exception of the Old Tampa Bay segment in 2009 and 2011,
chlorophyll a thresholds were not exceeded in all four major bay segments over the
2007-2011 period (TBEP. 2012a). The chlorophyll a threshold values are used as
indicators of impairment and vary slightly for each different segment of the bay, ranging
from 5.1 |ig/L for Lower Tampa Bay, to 8.5 |ig/L for Middle Tampa Bay, 9.3 |ig/L for
Old Tampa Bay, and 15 |ig/L for Hillsborough Bay [expressed as annual averages;
(TBEP. 2012aYI.
Seagrass Coverage
Five species of seagrasses, turtle grass (Thalassia testudinum), manatee grass
(Syringodium filiforme), shoal grass (Halodule wrightii), star grass (Halophila
engelmannii), and widgeon grass (Ruppia maritima), are found in Tampa Bay (TJSGS.
2011). These shallow-water grasses need sufficient light for photosynthesis and are
sensitive to reduced water clarity associated with increased algal growth. Scientists have
been tracking seagrass coverage as far back as the 1950s in Tampa Bay. Between the
1950s and 1980s seagrass coverage declined by about 50% or 9,000 ha in conjunction
with decreased water quality and physical impacts to the bay such as dredging and filling
(Poor etal.. 2013a; Haddad. 1989). Between 1999 and 2010, bay water clarity improved,
chlorophyll a concentrations decreased, and seagrass coverage area steadily increased,
coinciding with the reductions to N inputs to the bay rFigure 16-30; (TBEP. 2015; Poor et
al.. 2013a; TBEP. 2011; Greening and Janicki. 2006)1. The most recent surveys of
seagrasses (2014) exceeded the recovery goal set in 1995 [95% of estimated seagrass
coverage of the 1950s; (Sherwood et al.. 2016)1.
Through management actions, including an agreement with a power company to install
air pollution control systems and the completion of more than 250 projects to reduce N
inputs from the watershed, TN loadings to the bay have declined from previous annual
average estimates despite an ever-increasing population in the Tampa Bay metropolitan
area (Figure 16-31). Data compiled by the TBNMC indicate that continuing efforts to
reduce N loading are resulting in more-than-sufficient water quality for the expansion of
seagrasses (TBEP. 2012a).
16-102
DRAFT: Do Not Cite or Quote

-------
CO 18
O
"C—
x 16
CD
03 14 H
>
o
O 12
(/)
CO
l1
-------
2000 -I	T	1	T	,	T	|	T	,	.	,	T	,	T	,	T	1	.	,	T	,	T	,	T	,	T	,	.	1" 1.5
1984 1986 1988 1990 1992 1994 1996 1998 2000 2002 2004 2006 2008 2010 2012
Year
3750
-Hydrologically Normalized TN Loads
-Tampa Bay Metropolitan Area Population
2.75
3000
2750
O
2.25 -O
TN = total nitrogen; yr = year.
Source: TBEP (2012a).
Figure 16-31 Trend in hydrologically normalized total nitrogen load to Tampa
Bay relative to population increases in the Tampa Bay
metropolitan area.
1	Russell and Greening (2015) estimated a potential relative cost savings of $22 million per
2	year in avoided wastewater treatment plant costs due to nutrient reductions associated
3	with restoration and recovery of seagrass, marsh, and mangrove habitats in Tampa Bay.
4	Ecosystem services benefits of over $365 million estimated over an 18-year period from
5	1990 to 2008 include C sequestration and nutrient reductions via denitrification.
June 2018
16-104
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
16.4.4. Nitrogen Management
16.4.4.1. Numeric Nutrient Criteria
In 1998, the U.S. EPA published the National Strategy for the Development of Regional
Nutrient Criteria to promote the use of nutrient concentration levels in state water quality
standards (U.S. EPA. 1998b). Historically, Florida had a narrative nutrient water quality
criterion in place to protect waters against nutrient enrichment. In 2011, the state adopted
the first set of statewide numeric nutrient standards for Florida's waters. By 2015, almost
all of the remaining waters in Florida had numeric nutrient standards. Numeric nutrient
criteria in Florida are established for all estuary segments and include criteria for TN,
total phosphorus (TP), and chlorophyll a. For open ocean coastal waters, numeric criteria
are derived from satellite remote sensing techniques and established for chlorophyll a
only. Numeric nutrient criteria covering all of the mainstem segments and estuarine
segments in Tampa Bay were approved by the U.S. EPA in November 2012 (U.S. EPA.
2012c). A map of the bay segments is provided in Figure 16-23. The numeric nutrient
criteria for chlorophyll a for the mainstream segments of Tampa Bay are shown in
Table 16-19.
Table 16-19 Numeric nutrient criteria for chlorophyll a for the four mainstem
segments of Tampa Bay adopted by the Florida Department of
Environmental Protection.
Bay Segment
Chlorophyll a(|jg/L)
Old Tampa Bay
9.3 as annual mean
Hillsborough Bay
15.0 as annual mean
Middle Tampa Bay
8.5 as annual mean
Lower Tampa Bay
5.1 as annual mean
L = liter; |jg = microgram.
Modified from Sherwood et al. (2016).
Tampa Bay's numeric nutrient criteria for TN (Table 16-20) is expressed as tons/million
cubic meters of water as an annual total not to be exceeded more than once in a 3-year
period and represents an unorthodox approach to developing nitrogen nutrient criteria
(Florida DEP. 2016; Sherwood et al.. 2016). These values are based on previously
June 2018
16-105
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
adopted total maximum daily loads (TMDLs) or nutrient targets developed by TBEP.
Usually, the U.S. EPA's numeric nutrient criteria are expressed as a concentration, but
TMDLs are typically expressed as loads. The N management strategy for Tampa Bay
linked seagrass growth to external N loadings based on evidence that TN loads, resultant
chlorophyll and light attenuation, and seagrass response were closely correlated
(Sherwood et al.. 2016; Janicki Environmental. 2011). The TBNMC also found that when
freshwater inputs are greater, TN moves through the system more quickly; therefore,
residence time was an important consideration. TN loads and hydrologic loads affect both
the chlorophyll within the bay, and the amount of TN delivered per unit water was used
to set the numeric nutrient criteria for Tampa Bay. The FDEP and the U.S. EPA both
agreed that this framework, although unique for numeric nutrient criteria, would provide
reasonable assurance that the state water quality criteria would be met in Tampa Bay
(Sherwood et al.. 2016; Greening et al.. 2014).
Table 16-20 Numeric nutrient criteria for total nitrogen for the four mainstem
segments of Tampa Bay.
Segment
Nitrogen Delivery Ratio Threshold (tons/million m3)
Old Tampa Bay
1.08
Hillsborough Bay
1.62
Middle Tampa Bay
1.24
Lower Tampa Bay
0.97
m = meter; TBEP = Tampa Bay Estuary Program; TN = total nitrogen.
TBEP estuarine numeric nutrient criteria expresses as nitrogen delivery ratio (tons TN per million m3 hydrologic load) based on
1992-1994 conditions.
Source: Florida PEP (2016). Janicki Environmental (2011).
16.4.4.2. Nitrogen Management Consortium Efforts
The TBEP was established in 1992 as a part of the U.S. EPA's National Estuary Program
(Sherwood et al.. 2016). TBEP coordinates TBNMC, a public-private partnership
working to reduce N loading to the bay and whose members include 40+ local
governments (Tampa, St. Petersburg, and others) along with key industries bordering the
bay, such as electric utilities, fertilizer manufacturers, and agricultural operations.
June 2018
16-106
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
In recent years, one focus of the TBNMC's strategy has been to allocate N loads for
major point and nonpoint sources in order to meet state and federal requirements. The
governments and industries participating in the TBNMC voluntarily committed to cap
their N loads at average annual levels recorded in 2003-2007, which also meets the 1998
federally recognized TMDL for N. These capped allocations have been adopted by the
state of Florida and incorporated into NPDES discharge permits (TBEP. 2014). One
example of the Consortium's work is the reconfiguration of coal-fired plants in the
Tampa Bay watershed to reduce NOx emissions (Poor et al.. 2013a).
Every 5 years, the TBNMC must submit a Reasonable Assurance Update document to
FDEP for approval to document progress toward water quality and seagrass management
goals. The most recent reasonable Assurance Update (2012) presented data to show that
reasonable progress has been made towards the attainment of designated uses of Tampa
Bay. Based on this conclusion, FDEP upgraded Hillsborough Bay and Old Tampa Bay
segments to assessment Category 4b for nutrients (adequate management in place) from
Category 5 (impaired), and segments in Lower Tampa Bay were moved from
Category 4b to Category 2 (attains standards), signifying the state's recognition of
significant improvements in water quality and seagrass expansion (TBEP. 2014). The
Consortium members share the cost of the data collection and analysis to prepare the
Reasonable Assurance Update document, and the next update will cover progress made
during the 2013-2017 time period.
16.4.4.3. Response to Nitrogen Management Strategies
A recent paper reviewing the response of Tampa Bay to N management strategies
presented evidence of the ecosystem's recovery constituting a "regime shift" trajectory
(Duarte et al.. 2015; Duarte et al.. 2009). Since the mid 1980s Tampa Bay has shifted
back to a clear water seagrass system from a phytoplankton-dominated system that was
characteristic of the bay under previously eutrophic conditions. Following recovery from
an El Nino heavy rainfall period (1997-1998), water clarity in Tampa Bay increased
significantly, and seagrass expanded at a rate significantly different than before the event,
suggesting a feedback mechanism observed only in a few other systems. The authors
observed that too often these estuarine ecosystems are not able to recover even when
eutrophication stressors are mitigated because of the concurrent changes in environmental
conditions, that affected ecosystem dynamics occurring over the time period from the
onset of eutrophication to the reduction of nutrient levels. (Greening et al.. 2014) noted
four specific elements of the Tampa Bay management strategy that have resulted in a
successful recovery trajectory: (1) development of numeric water quality targets in
consultation with stakeholders, (2) citizen involvement in calling for action from
June 2018
16-107
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
regulatory agencies, (3) collaborative actions such as the TBNMC, and (4) state and
federal regulatory programs. BRACE scientists concluded that declines in atmospheric N
deposition to the estuary and its watershed were likely part of the historical improvement
in water clarity and increases in seagrass acreage (Poor et al.. 2013a; TBEP. 2011).
16.5. Rocky Mountain National Park Case Study
16.5.1. Background
This case study describes the environment and ecology of Rocky Mountain National Park
(ROMO) and summarizes research that provides insight into the impacts of atmospheric
nitrogen and sulfur deposition on terrestrial and aquatic ecosystems in ROMO. This
includes atmospheric deposition research from within ROMO, the surrounding area of the
Rocky Mountains, as well as other portions of the Northwestern Forested Mountains
ecoregion. Although the focus is on research published since 2008, publications prior to
2008 are used to establish context and identify long-term trends.
16.5.1.1. Description of Case Study Region
Rocky Mountain National Park (ROMO) was established by Congress in 1915 and is
located in the Front Range of the Rocky Mountains in north-central Colorado, about
80 km northwest of Denver. Straddling the Continental Divide, the western side of
ROMO serves as the headwaters for the Colorado River and eastern slope feeding the
South Platte River Basin (Wolfe et al.. 2003). The land within ROMO contains steep
elevation gradients and diverse ecosystem types, ranging from talus slopes at high
elevations and permanent glaciers above 4,000 m to grassy meadows at 2,400 m. Over
25% of ROMO is above the tree line (about 3,500 m), and ROMO contains large areas of
alpine tundra. In addition, there are more than 60 peaks above 3,600 m in elevation
(Porter and Johnson. 2007). Ecologically, ROMO is part of the Northwestern Forested
Mountains ecoregion (Figure 16-32).
June 2018
16-108
DRAFT: Do Not Cite or Quote

-------
Source: Clarke (2013V
Figure 16-32 Rocky Mountain National Park ecosystems.
June 2018
16-109
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
The ecoregion hosts extremely diverse vegetation and well-defined community zones
along elevation gradients. Winter snow cover averages 1.5 m annually in the Rocky
Mountains, and is a strong determinant of both plant and soil biodiversity (Clow ct al..
2016). More than 1,000 species of vascular plants have been documented in ROMO
(NPS. 2013). Within the alpine tundra, plant communities contain willow (Salix spp.) and
a variety of grasses, sedges, and forbs such as alpine sunflower (Rydbergia grandiflora)
and alpine forget-me-nots (Polemonium viscosum). The subalpine zone is composed
primarily of forests, dominated by Engelmann spruce (Picea engelmannii), subalpine fir
(Abies lasiocarpa), and limber pine (Pinus flexilis), whereas forests in the montane zone
vary from lodgepole pine (Pinus contorta) at cold high elevation sites, trembling aspen
(Populus tremuloides) in moist sites, and Douglas fir (Pseudotsuga menziesii) and
ponderosa pine (Pinus ponderosa) at more moderate and xeric sites, respectively
(Beidleman et al.. 2000).
More than 270 species of birds, including migratory species, have been reported in this
area over the last 100 years. Many are unique to mountainous habitats—aspen, ponderosa
pine, high-elevation willow, and spruce-fir vegetation types—found in the southern
Rocky Mountains. Seven native fish and four exotic fish inhabit the aquatic systems of
ROMO. Sixty-seven mammal species are known to be native to the area, but grizzly
bears (Ursus arctos), gray wolves (Canis lupus), and bison (Bison bison) were locally
extirpated in the 19th and early 20th centuries. The lynx (Lynx canadensis) and wolverine
(Gulo gulo) are either extirpated or extremely rare. Moose (Alces alces) are now
common, but were not historically recorded as part of this area of the Rocky Mountains.
A select group of four amphibian and two reptile species survive the high elevations and
cold temperatures in ROMO. They are all considered species of concern due to apparent
low numbers, lack of information about their status, and/or declining population trends.
There are 141 confirmed species of butterflies, but arthropod (insects, spiders, centipedes,
etc.) communities in ROMO have not been as well documented as other taxonomic
groups.
The NPS maintains a list of species known to occur in ROMO that are considered
endangered, threatened, or candidates for protection by the Endangered Species Act
(NPS. 2013):
•	Boreal toad (Bufo boreas boreas)—candidate
•	Yellow-billed cuckoo (Coccyzus americanus)—candidate
•	Canada lynx, (Lynx canadensis)—threatened
•	Greenback cutthroat trout (Oncorhynchus clarki stomias)—threatened
Pardo et al. (2011c) reported that responses to elevated N deposition in the Northwestern
Forested Mountains ecoregion include alteration of soil carbon:nitrogen (C:N) ratios,
16-110
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
base cation composition, and accelerated N cycling rates, including increased
mineralization and nitrification. Further changes described in this case study and in other
Appendices include plant, lichen, and algal chemistry; surface water chemistry (including
N concentration and acid neutralizing capacity); catchment N leaching rate; and changes
in the community composition of plants, lichens, and phytoplankton.
16.5.1.2. Class I Areas
Rocky Mountain National Park is designated as a federal Class I area. The Clean Air Act
(42 USC 7470) authorized these Class I areas to protect air quality in national parks over
6,000 acres and national wilderness areas over 5,000 acres in an effort to preserve pristine
atmospheric conditions. Class I areas are subject to the "prevention of significant
deterioration (PSD)" regulations under the Clean Air Act (42 USC 7470). New and
modified existing air pollution sources are required to have PSD preconstruction permits
and air regulatory agencies are required to notify federal land managers (FLMs) of any
PSD permit applications for facilities within 100 km of a Class I area.
16.5.1.3. Regional Land Use and Land Cover
The land area within the ROMO region falls predominantly into four land cover
classifications: evergreen forest, barren land, herbaceous cover, and perennial snow/ice
(Figure 16-33). The remainder of the land coverage categories account for less than 7%
of the total area. The area immediately surrounding ROMO is largely forested.
Regionally, the land to the north, west, and south of ROMO is predominantly forested,
with some grasslands used as grazing land. The region to the east and southeast contains
more grassland, agricultural, and urban land use. The smaller towns of Estes Park,
Allenspark, and Grand Lake are close to the boundaries of ROMO (Figure 16-34). while
the larger population centers of Fort Collins and the Denver metropolitan area are further
to the east and southeast, respectively (Figure 16-33 inset).
There are 29 total watersheds at the U.S. Geological Survey level-12 hydrological unit
code (HUC12s) scale that fall completely or partially inside ROMO (Figure 16-34).
June 2018
16-111
DRAFT: Do Not Cite or Quote

-------
Rocky Mountain National Park
Case Study Location
Land Cover Classification

•Evergreen Forest (52.7%)
	| "Barren Land (13,7%)
1 1
Hay/Pasture
I Deciduous Forest
1 1
"Herbaceuous (13.8%)
Developed, Low Intensity

Open Water
| Developed, Medium Intensity
i i
•Perennial Snow/Ice (12.6%)
| Developed, Open Space
i—in
Shrub/Scrub
Emergent Herbaceuous Wetlands
I I
Woody Wetlands
"Over 90% of coverage from these four
categories
m
Estes Park
m
Fort Collins
80 Miles
J a c k s o n
County
Routt National
Forest
Glen Haven
Grand Lake
COL
Jamestown
Boulder
Gr
Co u n
Aliens park
Lakewood
Denver
Figure 18-33 Rocky Mountain National Park land coverage using the land cover
classifications as mapped by the National Land Cover Dataset.
Percentage of cover is shown for the four dominant cover types.
June 2018
16-112
DRAFT: Do Not Cite or Quote

-------
Rocky Mountain National Park
Case Study Location
HUC12 Boundaries
Inside and outside of the park boundary
©
Allenspark
g Arapaho National
Forest
D Miles
Jackson
County
Grand Lake
HUC = hydrologic unit code.
Figure 16-34 Rocky Mountain National Park hydrologic unit code 12
watersheds.
June 2018
16-113
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
16.5.2.
Deposition
Based on TDEP calculations (see Appendix 2 of the ISA), most of ROMO is estimated to
receive deposition between 3-9 kg N/ha/yr and 3-6 kg S/ha/yr (Figure 16-35). In
contrast to the low rates observed throughout much of the western U.S., rates of N
deposition within ROMO are relatively high, and areas to the east of ROMO experience
deposition rates similar to areas in the north-central U.S. (see Figure 16-35 N deposition
inset). However, the estimated rates of S deposition are consistent with the low rates
observed elsewhere in the western U.S. and considerably lower than the S deposition
rates observed within the Ohio Valley. Within these TDEP estimates, the inorganic N
deposition in ROMO is relatively evenly balanced between oxidized and reduced forms,
with the northeastern portion of the region receiving predominantly reduced forms of N
(Figure 16-36). There are three National Atmospheric Deposition Program (NADP)
measurement sites within ROMO (Figure 16-36). including measurements stretching
back more than 30 years at the Loch Vale and Beaver Meadows sites. Over the past
25 years, sulfate (SO42 ) and hydrogen ion (H+) deposition have declined slightly at the
Beaver Meadows site (CO 19), whereas deposition of ammonium (NFLO and nitrate
(NO3 ) has been relatively constant (Figure 16-35). See Appendix 2.4. Appendix 25_ and
Appendix 2jS for more information on deposition in the U.S. Other maps showing the
contributions of individual species to dry and/or wet deposition are given in
Appendix 2.7.
Atmospheric N deposition generated from anthropogenic sources is a significant
influence on many ecosystems within ROMO (Wolfe et al.. 2003; Baron et al.. 2000;
Williams and Tonnessen. 2000; Williams et al.. 1996a; Caine. 1995). For this reason,
atmospheric N deposition has been the focus of considerable research, including the
Rocky Mountain Airborne Nitrogen and Sulfur (RoMANS) study (Beem et al.. 2010).
The RoMANS study was designed to characterize the sources as well as the transport,
transformation, and deposition processes of oxidized S and oxidized and reduced forms
ofN (Beem et al.. 2010).
Primary atmospheric transport to ROMO is from the west and northwest. However,
atmospheric circulation patterns along the Colorado Front Range can be complex, and
significant quantities of precipitation can fall on the east side of ROMO when easterly
upslope events occur (Porter and Johnson. 2007; Mast et al.. 2003; Baron and Denning.
1993). For instance, although prevailing pattern of atmospheric circulation brings
relatively clean air from west of ROMO, anthropogenic N can be transported to ROMO
from urban and agricultural areas along the Front Range when convective heating over
the plains produces easterly upslope winds (Figure 16-37). Higher than 3,000 m above
sea level (masl), westerly air flow dominates over the upslope pollution (Sievering et al..
June 2018
16-114
DRAFT: Do Not Cite or Quote

-------
1996). but coal-fired power plants located in Hayden and Craig, CO approximately
150 km west of RQMO, emit NOx that may reach ROMO with prevailing westerly
winds.
Annual Wet Deposition and 3-Year Moving Average at
Site C019: 1990-2014
200
180
160
hi 140
-C
"5 120
E
\ 100
c
,2 80
s «>
(X
OJ 40
Q
20
•		-	• T
1988	1992	1996	2000	2004
Year
2008
2012
2016
H+ = hydrogen ion; ha = hectare; kg =ki!ogram; mol = mole; N = nitrogen; NhV = ammonium; N03 = nitrate; RMNP = Rocky
Mountain National Park; S = sulfur; S042" = sulfate; yr = year.
Figure 16-35
Spatial patterns of atmospheric nitrogen and sulfur deposition in
the Rocky Mountain National Park region based on TDEP
calculations averaged from 2011-2013 (see Appendix 2) and
long-term trends in wet atmospheric deposition from the Beaver
Meadows National Atmospheric Deposition Program Monitoring
site within Rocky Mountain National Park.
June 2018
16-115
DRAFT: Do Not Cite or Quote

-------
~ Ouflv^r. CQ
O Monitor CO * 19
0 Monitor Location*
HMNF* Sludry Area
m
-	tri&
CoibCiKfo-
m,
%. 0»^.x*d W 0^pc-»ir<.n
¦¦mi r Hm
fJ*J*J* ** jf
~	D»flv*r,CO
O Moniflor CO * 19
#	Monitor Location*
Li RMNP Slucfy Area
m
0
_
Co4of»ab
% W«E Ckp-xi' .:-'.
//////
N = nitrogen; RMNP = Rocky Mountain National Park.
Figure 16-36
Spatial patterns in the percentage of atmospheric nitrogen
deposited in oxidized forms and through wet deposition in the
Rocky Mountain National Park region based on TDEP calculations
averaged from 2011-2013 (see Appendix 2).
June 2018
16-116
DRAFT: Do Not Cite or Quote

-------
IIKIShSl O'cLu
^^^litroger^ource^
chemical
conversion
and dispersion
C Transport I Transformation
(i
^^Dcposttio^^Foedbaci^TB
Particle and Gas
Transformations
Effects
^Vis|b|lit^ ^Otofw^ ^Cjjmate^Ownj^
Wet Deposition
Dry DcposHiOf
lightning
VOWWWI
ami olbor
area sources
ptents
urban and
sources
Surface and Ground
Water Pollution t-, ^ ^
*%SfdB
Ecosystem
Effects
Soil and
Natural
Vegetation
Fertilizer and
Feedlot Chemistry
Intro.
HN03 = nitric acid; NH3 = ammonia; NH4+ = ammonium; NO = nitric oxide; N02 = nitrogen dioxide; NCV = nitrate; N0X = NO + N02
Source: Clarke (2013).
Figure 16-37 Rocky Mountain National Park nitrogen cycle.
1	Numerous organic and inorganic N species contribute to total N deposition in ROMO.
2	Lee et al. (2016) estimated wet and dry deposition of reactive N species to ROMO using
3	a GEOS-Chem adjoint model, while Benedict et al. (2013) conducted a year-long survey
4	of the chemical speciation of N deposition at ROMO (Figure 16-38). Wet deposition of
5	inorganic N, particularly NH/, was the dominant form of N delivered to the surface at
6	ROMO. Wet deposition of organic nitrogen (ON) was also a major component of
7	deposition. Combined, reduced forms of N (wet and dry N H4 . dry NHj, ON) comprised
8	a large majority of the total N deposition flux in these measurements (Benedict et al..
9	2013).
June 2018
16-117
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14






Dry HNO.
Dry NU, V
Dry NO ¦
Dry PON ¦
Dry NO, W	










DO 0.2 0.4 0.6 OB 10 1.2 1.4
Total deposition (kg N/ha)
ha = hectare; HN03 = nitric acid; kg = kilogram; N = nitrogen; NH3 = ammonia; NH4+ = ammonium; N02 = nitrogen dioxide;
N03 = nitrate; ON = organic nitrogen; PON = particulate organic nitrogen.
Source: Benedict et al. (2013).
Figure 16-38 Total wet and dry deposition of nitrogen components measured at
Rocky Mountain National Park from November 2008 to November
2009, including organic nitrogen and particulate organic nitrogen.
Modeling conducted by Gebhart et al. (2011) suggests that the majority of the ammonia
measured in ROMO during the RoMANS study was emitted within Colorado. Modeling
conducted by (Lee et al.. 2016) suggest that reduced N deposition comes mostly from
sources east of the park, and oxidized N deposition comes primarily from sources west of
the park. Major emission sources of N to deposition in ROMO include ammonia from
livestock and oxidized N from mobile sources (Figure 16-39).
Aeolian dust deposition from the Colorado Plateau mitigates some of the acidifying
effects of S andN deposition in ROMO. Between 1993 and 2014, dust deposition
increased by as much as 81% in the southern Rocky Mountains. Clow et al. (2016)
estimated that deposition of dust has altered snowpack chemistry by increasing its
alkalinity (annual change of 0.32 |acq Ca2+/L/yr) over the same time period in which
smaller magnitude decreases in acid deposition to the snowpack (annual changes of
-0.10 |_ieq S0.f7L/yr and -0.07 |acq NO;, /L/yr) have occurred.
June 2018
16-118
DRAFT: Do Not Cite or Quote

-------
Emission sectors
NH3 (livestock)
[ NH3 (fertilizer)
^ NH3 (natural)
U NOx (surface inventory)
| NOx (electric generating units)
| NOx (non-electric generating
industrial stacks)
~| NOx (aircraft)
	NOx (lightning)
NOx (soil)
Geographic footprint of Nr deposition
» 0.03 0.(
[kg N/ha/yr]
RM (x2) 4.
Source: adapted from Figure 5 in Lee et al. (2016).
Figure 16-39 Annual-averaged monthly footprint of reactive N deposition in
Rocky Mountain National Park(4.0 kg N/ha/yr), and pie chart of
fractional contribution from emission sectors, as estimated by
GEOS-Chem adjoint model.
16.5.3.	Critical Loads and Other Dose-Response Relationships
1	This section focuses on studies that describe dose-response functions and/or critical loads
2	(CL) for atmospheric deposition effects on ecological endpoints. Terrestrial effects are
3	discussed first, followed by aquatic effects and a discussion on integration.
16.5.3.1. Terrestrial Critical Loads and Dose-Response Relationships
4	Among terrestrial ecosystems impacted by N deposition, alpine ecosystems are
5	particularly sensitive to increased availability of N due to inherently low rates of N
6	cycling, low rates of primary production, and thin, poorly weathered soils (Fenn et al..
7	1998). Among taxonomic groups in this ecoregion, the impact of atmospheric deposition
8	on epiphytic lichens is concerning because their strong dependence on atmospheric
June 2018
16-119
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
deposition for nutrients and unique symbiotic physiology makes them highly vulnerable
to N and S pollution and because lichens are important contributors to ecosystems
services (Brodo et al.. 2001). Among lichens, the structure and physiology of oligotrophic
species makes them especially important components of winter food webs, hydrologic
and nutrient cycles, and wildlife habitat (McCunc and Geiser. 1997).
16.5.3.1.1. Empirical Studies
In the Pardo et al. (201 lc) critical loads assessment, most of the studies examining the
responses of alpine ecosystems were conducted in the Colorado Front Range of the
Rocky Mountains, either in ROMO or in adjacent ecosystems such as the Niwot Ridge
Long-Term Ecological Research site. Critical load values for these alpine ecosystems
ranged from 1.2 to >20 kg N/ha/yr, excluding two values from Fenn etal. (2008) which
were determined for southwestern forested ecosystems in California (Table 16-21).
Table 16-21 Terrestrial empirical critical loads of nutrient nitrogen for the
Northwestern Forested Mountains ecoregion.
CL





Reliability

(kg N/



Deposition/

in Pardo et

ha/yr)
Species
Response
Method
Addition
Site
al. (2011c)
Reference
1.2 to
Epiphytic
Lichen
Application of
Total
Coastal
(#)
Geiser et al.
3.7
lichens
community
western
deposition: 0.8
Alaska

(2010)

(150 species)
change in
Oregon and
to





mixed-conifer
Washington
8.2 kg N/ha/yr





forests
model
(CMAQ)



2.5 to
Epiphytic
Lichen
Fenn:
Fenn:
Sierra
##
Fenn et al.
7.1
lichens
community
empirical CLs
Inorganic N
Nevada and

(2008)

Fenn: Letharia
change in
and DayCent
throughfall: 1.4
San

Geiser et al.

vulpina
mixed-conifer
modeling
to
Bernardino

(2010)

Geiser:
forests
Geiser:
71.1 kg N/ha/yr
Mountains



150 species

application of
Geiser: Total





western
deposition: 0.8






Oregon and
to 8.2 kg






Washington
N/ha/yr






model
(CMAQ)



June 2018
16-120
DRAFT: Do Not Cite or Quote

-------
Table 16-21 (Continued): Terrestrial empirical critical loads of nutrient nitrogen
for the Northwestern Forested Mountains ecoregion.
CL
(kg N/
ha/yr)
Species
Response
Method
Deposition/
Addition
Site
Reliability
in Pardo et
al. (2011c) Reference
Subalpine
forest
(Engelmann
spruce [Picea
engelmannii])
Increase in
organic horizon
N, foliar N,
potential net N
mineralization,
and soil
solution N,
initial increases
in N leaching
below the
organic layer
Baron:
Baron:
CENTURY
modeled total
model
deposition: 0.2
Rueth and
to 16.0 kg
Baron: field
N/ha/yr
sampling east
Rueth and
and west
Baron: total
slope forests
deposition: 3.2

to 5.5 kg

N/ha/yr (for

1992-1997)
N addition
Ambient:
experiment
6 kg N/ha/yr;

Addition: 20,

40, or

60 kg N/ha/yr
ROMO east
and west of
Continental
Divide
##
Baron et al.
(1994) Rueth
and Baron
(2002)
4 to 10 Alpine dry
meadow
(Carex spp.,
including Carex
rupestris)
Plant species
composition
change
Niwot
Ridge, CO
##
Bowman et
al. (2006)
Carex rupestris
Increase in
vegetation
cover
N addition
experiment
Total
deposition:
6 kg N/ha/yr
N additions:
20, 40, or
60 kg N/ha/yr
Niwot
Ridge, CO
##
Bowman et
al. (2006)
5 to 10
Ectomycorrhizal
Ectomycorrhizal
Expert
2002: bulk N
Kenai
fungi
fungi
judgment
deposition
Peninsula,
(-40 species)
community
extrapolated
across five
AK

structure
from marine
sampling sites


change in
west coast
ranged from


spruce (Picea)
spruce forest
0.15 to


forests

2.3 kg/ha/60
days
2008: Wet
deposition 2.8
to 7.9 kg
N/ha/yr

Alpine
Soil NOs"
N addition
Total
Niwot
terrestrial
leaching and N
fluxes

deposition:
6 kg N/ha/yr
N additions:
20, 40, or
60 kg N/ha/yr
Ridge, CO
(#)
Lilleskov
(1999)
Lilleskov et
al. (2001)
Lilleskov et
al. (2002)
Lilleskov et
al. (2008)
>20
Bowman et
al. (2006)
CL = critical load; CMAQ = Community Multiscale Air Quality model; ha = hectare; kg = kilogram; N = nitrogen; N03 = nitrate;
ROMO = Rocky Mountain National Park; yr = year.
Reliability rating: ## = highly reliable—a number of published papers show comparable results; # fairly reliable—the results of some
studies are comparable; (#) expert judgment—few empirical data are available, Critical Load based on expert judgment of those
ecosystems.
Source: (Pardo et al.. 2011c).
June 2018
16-121
DRAFT: Do Not Cite or Quote

-------
1	Since the publication of Pardo etal. (2011c) assessment, there have been several new
2	studies on terrestrial CLs (Table 16-22). New studies find CLs for soil N concentration
3	changes and leaching at 9.0-14.0 kg N/ha/yr, CLs to protect lichens at 4.0-4.1 kg
4	N/ha/yr, and CLs to protect vascular plant biodiversity (species richness, abundance, or
5	community composition) at 1.0-8.7 kg N/ha/yr.
June 2018
16-122
DRAFT: Do Not Cite or Quote

-------
Table 16-22 Montane forest and alpine ecosystem critical loads for nitrogen
deposition research published since the critical load assessment by
Pardo et al. (2011c).
CL (kg
N/ha/yr) Vegetation Response	Method Deposition/Addition Site	Reference
1.0
Composite
5% change in
ForSAFE-VEG
Deposition:
Northern and
SverdruD et al.

montane plant
alpine and
modeled
~4 kg N/ha/yr,
central
(2012)

communities
subalpine plant
change in
1 kg S/ha/yr
Rocky



community
future (to 2500)

Mountains



composition
vegetation

region

1.9
Salix Candida,
10% change in
ForSAFE-VEG
Deposition:
Loch Vale
McDonnell et al.

Carex spp.,
treeline
modeled
3.5 kg N/ha/yr
watershed,
(2014a)

Abies
(sub/alpine)
change in

ROMO


lasiocarpa,
plant
future (to 2100)




Geum rossii
community
vegetation





composition




3.0
Carex
Alpine
Field addition
Ambient deposition:
ROMO
Bowman et al.

rupestris
vegetation
study
4 kg N/ha/yr

(2012)


abundance

Addition: 5, 10, and






30 kg N/ha/yr.


4.0
Epiphytic
Degradation of
Empirical CLs
1.83 to
Northern
McMurrav et al.

lichens
lichen

3.45 kg N/ha/yr
Rocky
(2015)


communities

(CMAQ)
Mountains

4.1 Epiphytic Poorerthallus Empirical CLs Throughfall	Wind River McMurrav et al
lichens	condition	<0.9 to 4.1 kg N/ha/yr Range, WY, (2013)
L vulpina	including the
place I
U' ^ppon/ca	Brjdger
Wilderness
8.7
Forbs and
grasses—
open canopy
ecosystem
Species
richness loss
Empirical CLs
from vegetation
surveys
1 to 19 kg N/ha/yr,
wet (NADP) + dry
(CMAQ) deposition.
Over
15,000 plots
over the
continental
U.S.
Simkin et al. (2016)
9.0
Alpine dry
Soil solution
Field addition
Ambient deposition:
ROMO
Bowman et al.

meadow
NOs"
study
4 kg N/ha/yr

(2012)


concentrations

Addition: 5, 10, and






30 kg N/ha/yr


14.0
Alpine dry
meadow
Soil dissolved
inorganic N
concentrations
Field addition
study
Ambient deposition:
4 kg N/ha/yr
Addition: 5, 10, and
30 kg N/ha/yr
ROMO
Bowman et al.
(2012)
CL = critical load; CMAQ = Community Multiscale Air Quality model; ForSAFE-VEG = a dynamic forest ecosystem model; ha = hectare;
kg = kilogram; N = nitrogen; NADP = National Atmospheric Deposition Program; N03" = nitrate; ROMO = Rocky Mountain National
Park; S = sulfur; yr = year.
June 2018
16-123
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
Modeling Studies
Elevated lake and stream NOs concentrations have been observed in ROMO since the
mid 1980s. To understand the source of these elevated NO3 concentrations and the role
of atmospheric N deposition, Baron et al. (1994) used the CENTURY terrestrial
biogeochemistry model to estimate N uptake by plants and soils in alpine forest and
subalpine tundra within the Loch Vale watershed in ROMO. Based on the model output,
N uptake in the subalpine tundra was considerably lower than in the alpine forest.
Increased export of NO3 into stream water occurred at low N deposition rates (3 to
4 kg N/ha/yr) because much of this atmospheric N was deposited in the tundra or on
exposed rock surfaces, providing little N uptake. This work by Baron et al. (1994) was
the first N deposition CL research within the Northwestern Forested Mountains ecoregion
and aside from two CL estimates for lichens (Geiser et al.. 2010). the CL estimate
produced from this research is the lowest among the CL estimates reviewed by Pardo et
al. (2011c') for this ecoregion (Table 16-21).
Terrestrial biogeochemistry models have also played a role in the development of some
CL estimates for the ROMO region published since the Pardo et al. (2011c) assessment.
Sverdrup et al. (2012) and McDonnell et al. (2014a) used the ForSAFE-VEG model to
evaluate potential long-term effects of climate change and atmospheric N deposition on
alpine/subalpine ecosystems. Critical loads in both studies were defined as the rate of N
deposition at which plant diversity was protected from a change of 5 to 10 Mondrian
units (Mu; i.e., 5-10% change in plant species cover). Sverdrup et al. (2012) focused on a
"generalized" alpine/subalpine site, with vegetation composition constructed from plant
community data from ROMO and other national parks in the northern and central Rocky
Mountains region. Sverdrup et al. (2012) simulated plant responses to a future climate
(IPCC A2) and four levels of N and S deposition (preindustrial background, Clean Air
Act [CAA] controls, no CAA controls, and no CAA controls + high deposition). Soil
base saturation and soil solution base cation (Bc)-to-aluminum (Al) ratios (Bc:Al) were
predicted to decrease to less than 1% and less than 10% after Year 2100, respectively,
with the scenarios of no CAA emission controls and no CAA controls + high deposition
(indicating soil acidification). Future plant species coverages were predicted to change in
successively greater amounts in response to the altered climate, CAA emission controls,
no CAA emission controls, and no CAA emission controls + high deposition. Calculated
CLs (based on a change of 5 Mu) for N deposition were 1 kg N/ha/yr. Critical loads
related to S deposition were not discussed. All future N deposition scenarios (except
preindustrial background N) resulted in CL exceedance.
The approach of McDonnell et al. (2014a) was similar to that of Sverdrup et al. (2012)
because both applied the ForSAFE-VEG model to a subalpine ecosystem to determine
16-124
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
CLs for long-term vegetation composition shifts caused by climate change and
atmospheric N deposition. The former study differed from the latter by using the Loch
Vale watershed within ROMO as the study area and also back-projecting the model to
1900. McDonnell et al. (2014a) estimated that the N deposition CL to avert a 10% (5 Mu)
change in future (2010-2100) plant biodiversity was 1.9 to 3.5 kg N/ha/yr, which had
already been exceeded. Future air temperature increases were predicted to further change
plant community composition and exacerbate changes caused by N deposition alone. In
the simulations between 1900 and 2010, the authors found that subalpine fir (Abies
lasiocarpa) sapling coverage had increased by more than 25%, graminoid response was
mixed, and forbs generally had decreased in abundance (Table 16-23). By 2010, Geum
rossii was reduced by more than 50% of its simulated historical cover. In scenarios using
ambient N deposition and 0.5 times ambient N deposition, pronounced increases in the
abundance of the moss Aulacomnium palustre would occur in 2065 and 2080,
respectively, while the dominant graminoid (Carex rupestris) decreased. They also found
that higher total N deposition scenarios would suppress moss cover near the end of the
simulation period. Projected future tree responses were similar between the lowest two
and highest three total N deposition scenarios. Future forb and Carex rupestris cover
remained relatively constant under higher total N deposition rate scenarios.
16-125
DRAFT: Do Not Cite or Quote

-------
Table 16-23 Hindcast absolute and percentage changes in species abundance
between 1900 and 2010 in response to historical reconstructions of
nitrogen deposition (Sullivan et al.. 2005) and historical climate
change (IPCC. 2007b).
Relative Abundance %
Growth Form
Species
1900
2010
Absolute
Change
% Change
Tree
Abies lasiocarpa
6.3
8.4
2.1
33.3
Shrubs
Salix Candida
15.4
25.5
10.1
65.6

Rubus parviflorus
0.0
4.4
4.4
—
Graminoids
Carex rupestris
6.8
16.4
9.6
141.2

Carex elynoides
25.1
12.8
-12.3
-49.0

Festuca ovina
1.7
3.2
1.5
88.2

Calamagrostis purpurascens
4.3
3.0
-1.4
-32.6

Poa abbreviata
3.2
2.3
-0.9
-28.1
Forbs
Geum rossii
19.7
8.9
-10.8
-54.8

Aquilegia caerulea
4.6
4.8
0.2
4.3

Antennaria rosea
1.5
2.0
0.5
33.3

Arenaria fendleri
4.7
1.8
-2.9
-61.7

Minuartia obtusiloba
3.7
1.6
-2.0
-54.1

Zigadenus elegans
0.0
2.2
2.2
—
Moss
Aulacomnium palustre
3.0
2.6
-0.5
-16.7
Source: McDonnell et al. (2014a).
16.5.3.2.	Aquatic Critical Loads
1	Nitrogen limitation of algal productivity is widespread in remote and undisturbed
2	freshwater lakes, such as the high-elevation lakes located throughout the Rocky
3	Mountains (Baron et al.. 201 lb). Based on surveys conducted in the 1980s that quantified
4	the ratio of dissolved inorganic N to total phosphorus, alpine lakes in the Rocky
June 2018
16-126
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
Mountains region are predominantly N limited (45% of lakes) or colimited by both N and
P [22% of lakes; (Pardo et al.. 201 IcYI. This may underestimate the previous extent of N
limitation in these systems because lake sediment records indicate that N deposition
began impacting biogeochemistry in some alpine lakes during the 1950s (Das et al..
2005; Wolfe et al.. 2003; Wolfe et al.. 2001). Pardo et al. (2011c) summarized CLs for
ROMO surface waters (Table 16-24) with empirical and modeled CL estimates ranging
from 1.5 to 2.5 kg N/ha/yr. Within ROMO, eastern slope surface waters commonly have
higher dissolved inorganic N (DIN) than western slope surface waters (Wolfe et al..
2001; Baron et al.. 2000).
While the majority of ROMO aquatic research since 2000 focused on eutrophication,
Vertucci and Corn (1996) and Wolfe et al. (2003) described the potential for aquatic
acidification in the Colorado Rocky Mountains. Neither Vertucci and Corn (1996) nor
Wolfe et al. (2003) found evidence for acidification, citing the influence of catchment
sources of base cation neutralization. However, Wolfe et al. (2003) suggested that
persistent N deposition threatened to cause chronic surface water acidification,
particularly given that episodic declines of pH and acid neutralizing capacity (ANC) have
been well documented in east slope headwaters of the Front Range (Williams and
Tonnessen. 2000; Williams et al.. 1996b). Vertucci and Corn (1996) found that although
episodic acidification occurred during snowmelt, at no point was ANC < 0 in the northern
Colorado and southern Wyoming lakes and streams that were sampled. Importantly, the
observed acidification episodes did not coincide with the presence of amphibian embryos,
meaning that amphibians in the region were not threatened by acidification (including the
boreal toad, a candidate species for Endangered Species Act protection).
16-127
DRAFT: Do Not Cite or Quote

-------
Table 16-24 Critical loads for nitrogen for eutrophication for surface water
(high-elevation lakes) in the Rocky Mountains.
CL kg



Deposition/


N/ha/yr
Species
Response
Method
Addition
Site
Reference
1.5
Diatom
Eutrophication
Used exponential equations
Wet deposition
Southern
Bowman

community

correlating with NOx
2.94 kg N/ha/yr
Rockies/
et al.

composition

emissions from CO and
Dry deposition
Loch Vale
(2006)



11 western U.S. states to
0.94 kg N/ha/yr
ROMO




reconstruct historic N





deposition. 1950-1964 wet






N deposition correlated with






alteration of ROMO diatom






assemblages.



1.5
F. crotonesis
Eutrophication
Paleolimnological—analyzed
Not specified
Northern
Saros et

A. formosa

diatom fossil records of four

Rockies/
al. (2003)



lakes in the Beartooth

Beartooth




Mountains. Small Fragilaria

Mountains,




species declined while

WY




Fragilaria crotonensis,






Asterionella formosa, and






multiple Cyclotella spp.






increased



2.0
Not
Eutrophication
Used CENTURY model to
Modeled total
Southern
Baron et

applicable

discern if high lake and
deposition of 0.2
Rockies/
al. (1994)



stream N measurements are
to 16.0 kg
Loch Vale




attributed to alpine tundra
N/ha/yr
ROMO




and subalpine forest N






saturation



2.5
Chlorophyll a
Eutrophication,
Compared oligotrophic lake
Wet DIN
Rocky
Berqstrom


N and P
chemical and chlorophyll a
deposition 2.5 to
Mountains
and


colimitation
data in 42 European and
3.5 kg N/ha/yr for

Jansson



North American regions to
regions 26 to 29

(2006)



inorganic N deposition data
(Rocky






Mountains in






area of ROMO)


3.0
Not
Increases in
Compared observations of
Modeled total
Rocky
Baron et

applicable
lake nitrate
nitrate concentrations in
deposition of 1.5
Mountains
al. (2011b)


concentration
285 lakes to observations to
to 7.5 kg N/ha/yr





N deposition estimates






derived from NADP and






PRISM



CL = critical load; CO = carbon monoxide; DIN = dissolved inorganic nitrogen; ha = hectare; kg = kilogram; N = nitrogen;
NADP = National Atmospheric Deposition Program; N03" = nitrate; NOx = NO + N02; P = phosphorus; PRISM = a model for
spatial climate data; ROMO = Rocky Mountain National Park; yr = year.
Source: Baron et al. (2011 by Pardo et al. (20110). Williams and Tonnessen (2000).
June 2018
16-128
DRAFT: Do Not Cite or Quote

-------
16.5.3.2.1. Empirical Studies
1	Pardo etal. (2011c) compiled a list of N deposition studies in high-elevation lakes that
2	have observed changes in the growth, biomass, or composition of phytoplankton
3	(Table 16-25). Within these studies, there was a relatively narrow range of lake water
4	nitrate concentrations at which these phytoplankton responses were observed
5	(Table 16-25). In paleolimnological studies of alpine lakes in this region that experienced
6	increased rates of deposition, there were pronounced shifts in the composition and
7	productivity of phytoplankton communities that occurred around 1950 (Table 16-26).
Table 16-25 Lake water nitrate concentrations in nitrogen deposition studies
observing phytoplankton responses.
N Cone,
(mg NOs'-N/L)
Species
Response
Deposition/Addition
Site
Reference
0.9
A. formosa, F.
crotonensis
Stimulation of
A. formosa, F.
crotonensis
Wet deposition
<2 kg N/ha/yr; N addition of
18 pmol N as NaNOs;
N + P enrichment, 18 pmol
N + 5 pmol P as NahtePCM
Beartooth
Mountains, WY
Saros et al.
(2005)
0.9
Chlorophyll a
Increasing
biomass
Wet DIN deposition 1.3 to
11 kg N/ha/yr; total N
deposition <0.1 kg N/ha/yr
to >15 kg N/ha/yr
Sweden
Berqstrom et al.
(2005)
0.9
Chlorophyll a
Increasing
biomass
Wet DIN deposition 2.5 to
3.5 kg N/ha/yr for regions
26 to 29 (Rocky Mountains
in an area of ROMO)
European and N.
American lakes,
including Rocky
Mountains
Berqstrom and
Jansson (2006)
1.0
Chlorophyll a
Shift in species
composition
Ambient deposition not
specified; single pulse N
treatment of 1,000 pg/L
NO3--N as KNOs
Snowy Range,
WY
Nvdick et al.
(2004)
1.1
Chlorophyll a
Diatom spp.
growth
stimulation
Ambient deposition of
3.5 kg N/ha/yr; N treatment
of ambient N03"-N plus
1 mg N03"-N/L
Colorado Front
Range (Loch
Vale)
Lafrancois et al.
(2003a)
Cone. = concentration; DIN = dissolved inorganic nitrogen; ha = hectare; kg = kilogram; KN03 = potassium nitrate; L = liter;
|jg = microgram; mg = milligram; N = nitrogen; NaH2P04 = monosodium phosphate; NaN03 = sodium nitrate; N03"-N = nitrate N;
P = phosphorus; ROMO = Rocky Mountain National Park; yr = year.
Source: Pardo et al. (20110).
June 2018
16-129
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
Table 16-26 Paleolimnological biological responses in Rocky Mountain lakes
exposed to anthropogenic nitrogen deposition.
Species
Response
Deposition/
Addition
Site
Year(s) of
Shift
Reference
A. formosa
Increased microbial activity,
greater primary production,
change in species
composition
Not specified
Colorado Rocky
Mountains
1950
Enders et al.
(2008)
A. formosa, F.
crotonensis
Greater primary production,
stimulation of A formosa
and F. crotonensis, shift in
species assemblages
Inorganic N
deposition
4 kg N/ha/yr
Eastside ROMO
lakes compared to
remote Colorado
Rocky Mountains
wilderness lake
1950
Das et al.
(2005)
A. formosa, F.
crotonensis
Greater primary production,
stimulation of A formosa
and F. crotonensis, shift in
species assemblages
Wet inorganic N
deposition 2 to
4 kg N/ha/yr
Dry deposition:
0.9 kg N/ha/yr
Colorado Rocky
Mountains
1950
Wolfe et al.
(2001). Wolfe
et al. (2003)
A. formosa, F.
crotonensis
Greater primary production,
stimulation of A formosa
and F. crotonensis, shift in
species assemblages after
1995
Not specified
Beartooth
Mountain Range,
WY (Yellowstone
NP region)
1995
Saros et al.
(2003)
ha = hectare; kg = kilogram; N = nitrogen; NP = national park; ROMO = Rocky Mountain National Park; yr = year.
Source: Pardo et al. (20110).
Since the Pardo et al. (2011c) CLs synthesis, Baron etal. (2011b) synthesized aquatic N
cycling research for lakes in several regions of the U.S., including alpine lakes in the
Rocky Mountains. Using existing water chemistry measurements and modeled estimates
of wet and dry N deposition, Baron etal. (2011b) observed that lake nitrate
concentrations increased where N deposition exceeded 2.0 kg N/ha/yr. This threshold
was similar to previous CLs of 1.5-4 kg N/ha/yr for lakes in this region (Table 16-24).
Baron etal. (2011b) also sought to develop a CL for N driven episodic aquatic
acidification, but found little data apart from that used to generate the (Williams and
Tonnessen. 2000) CL estimate of 4.0 kg N/ha/yr. Using phytoplankton biomass N to P
limitation shifts as the basis for CL calculations, Williams et al. (2017b) determined an
empirical CL of 4.1 kg/TN/ha/yr for remote high-elevation lakes across the western U.S.
The CLs were calculated as the total (wet + dry) N deposition rate at which point below
the CL, N limitation is more likely than P limitation and above the CL, P limitation is
more likely than N limitation. DIN:TP and DIN response categories yielded an average
critical load of 3.8 kg/TN/ha/yr for the lakes.
June 2018
16-130
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
Modeling Studies
In addition to the empirical estimates of aquatic CLs, Sullivan et al. (2005) and Hartman
et al. (2007) modeled CL estimates for aquatic acidification in ROMO. Sullivan et al.
(2005) used the MAGIC model to evaluate the sensitivity of two water bodies in the Loch
Vale watershed within ROMO—the Loch and Andrews Creek—to N and S deposition
based on ANC. Hartman et al. (2007) also modeled deposition scenarios for changes in
ANC at Andrews Creek, but used the DayCent-Chem model. Beginning with 1996
deposition rates of 2.2 kg S/ha/yr and 4.2 kg N/ha/yr, Sullivan et al. (2005) estimated CLs
for N or S in order to reach ANC of 0, 20, or 50 (ieq/L by 2046 in the Loch and Andrews
Creek, which had 1996 ANC values of 53 and 34, respectively. Acidification of the
watershed to ANC to 0 or 20 by 2046 in either water body would require doubling of
either N or S deposition (Sullivan et al.. 2005). However, increasing ANC to 50 in
Andrews Creek was not achievable by 2046 in any deposition scenario considered by
(Sullivan et al.. 2005). Hartman et al. (2007) took a somewhat different approach,
modeling 47-year scenarios beginning in 2000 where N deposition increased at rates that
were low (+60%), moderate (+120%), and high (+240%) relative to the contemporary
deposition rate. In these scenarios, episodic acidification (ANC = 0 (j,eq/L) of Andrews
Creek occurred at 6.3-7.1 kg N/ha/yr. Estimates for persistent (chronic) changes in ANC
of Andrews Creek were similar to those of Sullivan et al. (2005). within 1 kg N/ha/yr for
ANC = 20 |icq/L and within 3 kg N/ha/yr for ANC = 0 |icq/L (Table 16-27). Notably, the
Loch does not represent the most acid-sensitive lakes in the Front Range, with
approximately l/5th of the alpine lakes in the Colorado Front Range having a similar or
lower ANC (Eilers et al.. 1989). Williams et al. (2017b) modeled CLS for remote
high-elevation lakes across the Western U.S. based on nutrient limitation shift thresholds.
Modeled critical loads ranged from 2.8 to 5.2 kg/TN/ha/yr and correctly predicted
exceedances in 69% of lakes using NOs -N.
16-131
DRAFT: Do Not Cite or Quote

-------
Table 16-27 Critical loads of nitrogen or sulfur for surface water acidification
Rocky Mountain National Park and other high-elevation lakes in the
Rocky Mountains.
CL kg/ha/yr
of N or S
Species Response
Method
Deposition/
Addition
Site
Reference
4.0 (N)
Not
applicable
Episodic
freshwater
acidification
(ANC < 0 pmol/L)
1995 synoptic survey of NADP DIN wetfall Central
9 high-elevation lakes:
headwater catchments
experienced episodic
acidification due to
inorganic N in wetfall
(ANC < 0 pmol/L in
surface waters during
snowmelt).
at 23 sites.
>2,500 masl: 2.5
to 3.5 kg N/ha/yr
<2,500 masl:
<2.5 kg N/ha/yr
Rockies/
Colorado
Front
Range
Williams
and
Tonnessen
(2000)
5.8 (N)	Not	Freshwater
applicable acidification
(ANC = 50 peq/L)
MAGIC model scenario
for 1996 to 2046
Deposition:
2.2 kg S/ha/yr,
4.2 kg N/ha/yr
The Loch
(ROMO)
Sullivan et
al. (2005)
6.3 to 7.1 (N)
Not
Episodic
DayCent-Chem model
Deposition:
Andrews
Hartman et

applicable
freshwater
for 2000 to 2047 of four
2.7 kg S/ha/yr,
Creek
al. (2007)


acidification
N deposition (current,
3.5 kg N/ha/yr
(ROMO)



(ANC < 0 peq/L)
+60% increase, +120%






increase, +240%






increase)



7.1 (N)
Not
Freshwater
DayCent-Chem model
Deposition:
Andrews
Hartman et

applicable
acidification
for 2000 to 2047 of four
2.7 kg S/ha/yr,
Creek
al. (2007)


(ANC = 20 peq/L)
N deposition (current,
3.5 kg N/ha/yr
(ROMO)




+60% increase, +120%






increase, +240%






increase)



7.8 (N)
Not
Freshwater
MAGIC model scenario
Deposition:
Andrews
Sullivan et

applicable
acidification
for 1996 to 2046
2.2 kg S/ha/yr,
Creek
al. (2005)


(ANC = 20 peq/L)

4.2 kg N/ha/yr
(ROMO)

12.2 (N)
Not
Freshwater
MAGIC model scenario
Deposition:
Andrews
Sullivan et

applicable
acidification
for 1996 to 2046
2.2 kg S/ha/yr,
Creek
al. (2005)


(ANC = 0 peq/L)

4.2 kg N/ha/yr
(ROMO)

14.7 (N)
Not
Freshwater
MAGIC model scenario
Deposition:
The Loch
Sullivan et

applicable
acidification
for 1996 to 2046
2.2 kg S/ha/yr,
(ROMO)
al. (2005)


(ANC = 20 peq/L)

4.2 kg N/ha/yr


14.6 to 15.3
Not
Chronic
DayCent-Chem model
Deposition:
Andrews
Hartman et
(N)
applicable
freshwater
for 2000 to 2047 of four
2.7 kg S/ha/yr,
Creek
al. (2007)


acidification
N deposition (current,
3.5 kg N/ha/yr
(ROMO)



(ANC < 0 peq/L)
+60% increase, +120%






increase, +240%






increase)



20.6 (N) Not	Freshwater	MAGIC model scenario Deposition:	The Loch Sullivan et
applicable acidification for 1996 to 2046	2.2 kg S/ha/yr, (ROMO) al. (2005)
(ANC = 0 peq/L)	4.2 kg N/ha/yr
June 2018
16-132
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
Table 16-27 (Continued): Critical loads of nitrogen or sulfur for surface water
acidification Rocky Mountain National Park and other
high-elevation lakes in the Rocky Mountains.
CL kg/ha/yr
of N or S
Species
Response
Method
Deposition/
Addition
Site
Reference
2.8 (S)
Not
applicable
Freshwater
acidification
(ANC = 50 peq/L)
AGIC model scenario
for 1996 to 2046
Deposition:
2.2 kg S/ha/yr,
4.2 kg N/ha/yr
The Loch
(ROMO)
Sullivan et
al. (2005)
4.6 (S)
Not
applicable
Freshwater
acidification
(ANC = 20 peq/L)
MAGIC model scenario
for 1996 to 2046
Deposition:
2.2 kg S/ha/yr,
4.2 kg N/ha/yr
Andrews
Creek
(ROMO)
Sullivan et
al. (2005)
7.8 (S)
Not
applicable
Freshwater
acidification
(ANC = 20 peq/L)
MAGIC model scenario
for 1996 to 2046
Deposition:
2.2 kg S/ha/yr,
4.2 kg N/ha/yr
The Loch
(ROMO)
Sullivan et
al. (2005)
8.1 (S)
Not
applicable
Freshwater
acidification
(ANC = 0 peq/L)
MAGIC model scenario
for 1996 to 2046
Deposition:
2.2 kg S/ha/yr,
4.2 kg N/ha/yr
Andrews
Creek
(ROMO)
Sullivan et
al. (2005)
11.1 (S)
Not
applicable
Freshwater
acidification
(ANC = 0 peq/L)
MAGIC model scenario
for 1996 to 2046
Deposition:
2.2 kg S/ha/yr,
4.2 kg N/ha/yr
The Loch
(ROMO)
Sullivan et
al. (2005)
ANC = acid neutralizing capacity; CL = critical load; DIN = dissolved inorganic nitrogen; ha = hectare; kg = kilogram; L = liter;
|jeq = microequivalent; |jmol = micromole; MAGIC = a biogeochemical process model; masl = meters above sea level;
N = nitrogen; NADP = National Atmospheric Deposition Program; ROMO = Rocky Mountain National Park; S = sulfur; yr = year.
16.5.3.3. Integration
Together, there is a considerable amount of research available on the effects of
atmospheric N deposition on ecological processes and characteristics within ROMO.
Assembled together, the CL estimates developed in and around ROMO (Figure 16-40)
show a continuum of effects as atmospheric N loads increase from natural background
rates (0.2 kg N/ha/yr) to levels several times greater than the current estimates of
3-9 kg N/ha/yr (Figure 16-35). Broadly, the chemistry and biota of aquatic systems
appear to be the most sensitive to N deposition, followed by alpine plant communities
(Figure 16-39). It is notable that many of the published CLs (Table 16-21. Table 16-22.
and Table 16-24) are near or below the rates of atmospheric N deposition currently
observed in ROMO. Thus, there is considerable evidence that current rates of
atmospheric N deposition are influencing the composition and function of aquatic and
terrestrial ecosystems within ROMO [Figure 16-41; (RMNP Initiative. 2014; Porter and
Johnson. 2007)1
June 2018
16-133
DRAFT: Do Not Cite or Quote

-------

Rocky Mountain alpine lakes: shift in diatom community dominance.
. Baron etal., 2006 (ROMO) (1.5)
. Baron et a!., 2011 (285 RM lakes) (1.0)
Saros et al., 2005 (other RM) (1.5)
. Baron et al., So. (RM Loch Vale ROMO) (1.5)
Diatom assemblages already dominated by species indicative of moderate N enrichment, N limited oligotrophic
lakes switch to P-limitation after receiving only modest inputs of reactive N shifting the controls on diatom
species changes along the length of the nitrate gradient
Arnett et al., 2012 46 high elevation Rocky Mountains lakes (1-3.2 wet deposition)
Chlorophyll a concentration increase
• Elser et al., 2009a Colorado eastern lakes (2-7)
Eutrophication and P co-limitation lakes
• Bergstrom & Jansson, 2006 (RM lake) (2.5)
Diatom growth in A. formosa
Nanus et al., 2012 (location?) (>3)
Alpine vegetation
Bowman et al., 2011 (ROMO) (3.0)
Increased foliar chemistry, mineralization, nitrification, and nitrate leaching
Reuth & Baron, 2002; Reuth et al., 2003, ROMO (<4)
Alpine vegetation change (Corex rupestris indicator)
Bowman et al, 2006 Niwot Ridge (4.0)
Episodic acidification of surface waters
Baron et al., 2011 Niwot Ridge (4.0)
Williams and Tonnesson, 2000 Central Rockies, Colorado Front
Range
RR-N:RR-P ranging from 0.33 (P limitation) to 1.16 (N-P-co-
limitation) with the exception of 1 lake outlier of 1.42 (Green
Lake)
. Elser etal., 2009b (>6)
Nitrate leaching below rooting zone in alpine
Bowman et al., 2011, ROMO
I Community changes in alpine vegetation
| . Bowman et al, 2006 South Rockies Niwot Ridge (10.0)
Nitrate increases in soil solution in alpine
Bowman et al., 2011 ROMO (14.0)
O	5	10	15	20
Nitrogen deposition (kg/ha-yr)
ha = hectare; kg = kilogram; N = nitrogen; P = phosphorus; RM = Rocky Mountain; ROMO = Rocky Mountain National Park;
RR = relative response to changes in N or P; yr = year.
Figure 16-40 The continuum of ecological sensitivity to nitrogen deposition.
June 2018	16-134	DRAFT: Do Not Cite or Quote

-------
Chronic
acidification
Episodic
acidification
o
fi
!fc
LU
E
Ł
Vi
>s
W
o
o
LU
Change in alpine
plant species
Changes in soil
and foliar
chemistry
Change in
phytoplankton
composition and
abundance
Increased lako
and stream nitrate
J Potontial Futuro Effects
Current Effects

p
m
Ł3

N Load (kg-ha'yr1)
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Notes: current rates of N deposition impact terrestrial and aquatic ecosystems and relatively small increases in deposition may
cause further ecological change.
Source: Porter and Johnson (2007).
Figure 16-41
Critical load thresholds for current and possible future
biogeochemical and biological effects of nitrogen deposition.
16.5.4. Highlights of Additional Research Literature and Federal Reports since
January 2008
A number of studies documenting the effects of N or S deposition on terrestrial and
aquatic systems in ROMO or the ROMO region have been published since 2008. We
discuss these studies briefly here, but refer readers to the appropriate portion of the main
body of the ISA for a more detailed review of this work.
16.5.4.1. Terrestrial
A number of the studies reviewed in Appendix 4 and Appendix 6 of the current NOx-SOx
ISA as part of the body of terrestrial N deposition research published since 2008 were
conducted in and around ROMO. Farrer et al. (2013) and Bowman et al. (2012) both
conducted N addition studies in alpine tundra ecosystems to understand the response of
June 2018
16-135
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
plants to N deposition. Farrer et al. (2013) did not find an increase in net primary
productivity, but found both positive and negative effects on the aboveground biomass of
individual plant species. Bowman et al. (2012) did not find any significant overall effect
on aboveground plant biomass, plant tissue N concentration, or plant diversity; but the
abundance of an important sedge (Carex) species almost quadrupled. This contrasts with
Farrer et al. (2015) and Yuan et al. (2016). who found that N additions decreased plant
species diversity. In mixed grass meadows near ROMO, an experiment that recreated
historical, low N deposition conditions (soil N availability reduced by 63%, winter
precipitation reduced 50%) reduced biomass of invasive cool-season grass Bromus
tectorum (Concilio et al.. 2016).
There are no published CLs for changes in soil microbial communities in the ROMO
area, but several studies in alpine tundra ecosystems have quantified microbial responses
to N additions ranging from 8 to 288 kg N/ha/yr. Dean et al. (2014) observed decreased
ericoid fungi abundance and decreased richness and diversity of root-associated fungi,
while Farrer etal. (2015) and Boot et al. (2016) observed decreased microbial biomass,
fungal biomass, and bacterial biomass. Farrer et al. (2013). Nemergut et al. (2008). and
Yuan etal. (2016) documented shifts in microbial community composition.
Lieb et al. (2011) found that a decade of N additions to alpine ecosystems lowered soil
acid buffering capacity, decreased concentrations of base cation Mg2+, and increased
concentrations of Mn2+ and Al3+. After 15 years of N additions, Boot et al. (2016) found
that N addition reduced soil pH, soil percentage of C, and soil C:N in the O-horizon.
16.5.4.2. Aquatic
As with terrestrial ecosystems, there have been a number of new studies quantifying the
effects of N deposition on aquatic ecosystems in the Rocky Mountains since 2008
(Table 16-28). McCrackin and Elser (2012) measured denitrification in sediments of
alpine lakes in the Colorado Rocky Mountains that received elevated (5-8 kg N/ha/yr) or
low (<2 kg N/ha/yr) N deposition inputs. In high-deposition lakes, the NOs -N
concentration was significantly higher than in low-deposition lakes, but there was also
evidence that denitrifying bacteria could remove a meaningful portion of N inputs to the
lakes. While current levels of N deposition have not saturated the microbial capacity for
denitrification, the abundance of denitrifying bacteria has not increased in response to
additional N.
June 2018
16-136
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
Table 16-28 Summary of freshwater eutrophication studies in the Rocky
Mountains since 2008.
Species
Response
Method
Deposition/
Addition
Site
Reference
Not	Diatom assemblages
applicable already dominated by
species indicative of
moderate N enrichment
Diatom calibration
from surface
sediment samples
Wet deposition: 1 to
3.2 kg N/ha/yr
46
Arnett et al.
high-elevation (2012)
lakes in the
Rocky
Mountains
Not	Average lake DIN:TP of	Mesocosms	Ambient deposition Colorado Gardner et
applicable 16.3 + 2.76, suggesting	amended with	not specified	Front Range al. (2008)
widespread N saturation,	NO3", PO43", or	Addition'
but unclear if this is a new	PO43" + NO3"	930 mg/L NO3"
or historical condition
Chlorophyll a Greater N deposition
increased the N:P ratio in
lakes. Phytoplankton
growth P limited in high N
deposition lakes
(~7 kg N/ha/yr), but N
limited in low-N deposition
lakes (~2 kg N/ha/yr)
N and P enrichment
bioassay
experiments in
high- and
low-deposition
lakes
Deposition: 2 to
7 kg N/ha/yr
14 eastern
Colorado
alpine lakes
Elser et al.
(2009a)
Chlorophyll a
Chi a conc. increase.
Responses indicated a
range from P limitation to
N-P colimitation with one
N limited outlier lake.
Enrichment
bioassay
Wet deposition
>6 kg N/ha/yr
14 eastern
Colorado
alpine lakes
Elser et al.
(2009b)
A. formosa Maximum diatom growth Growth kinetics Total deposition Rocky	Nanus et al.
rate at 0.5 |jM of NO3" experiments	>3 kg N/ha/yr	Mountains (2012)
Chi a = chlorophyll a; DIN = dissolved inorganic nitrogen; ha = hectare; kg = kilogram; L = liter; |jM = micromolar; mg = milligram;
N = nitrogen; N03" = nitrate; P = phosphorus; P043" = phosphate; TP = total phosphorus; yr = year.
Mastetal. (2011) examined trends in precipitation chemistry and other factors that
influence long-term changes in chemistry of high-elevation lakes in three Colorado
wilderness areas. Mastetal. (2011) observed that sulfate concentrations in precipitation
decreased at rates of-0.15 to -0.55 (j,eq/L/year during the Years 1985 through 2008 at
10 monitoring stations in Colorado. In high-elevation lakes where SO42 was primarily
derived from atmospheric sources, the lake SO42 concentrations decreased by -0.12 to
-0.27 (ieq/L/year. In lakes where watershed sources were the dominant source of SO42 .
the S042 concentrations in lake water increased during this time period. Mast et al.
(2011) attributed this trend to the effects of climate warming on pyrite weathering.
Similarly, Heath and Baron (2014) observed increases in summer (June-August) fluxes
and concentrations of cations, SiC>2, SO42 . inorganic N, and ANC in stream water from
June 2018
16-137
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
1984-2010 within the Loch Vale watershed in ROMO and attributed these trends to
climate change effects on nitrification, mineral weathering, and the release of elements
entrained in snow and ice. In comparison, wet N deposition rates were relatively constant
and wet SOr deposition decreased.
Mast et al. (2014) measured long-term changes in stream NOs concentration over three
decades at the Loch Vale watershed in RMNP. The concentrations of NO3 in stream
water increased during the early 1990s, peaked in the mid 2000s, and then declined by
more than 40%. Nanus et al. (2012) developed a surface water NO3 threshold for growth
of the diatom A. formosa for high-elevation lakes in the Rocky Mountains. Arnett et al.
(2012) tried to develop a CL for diatom community composition change along a gradient
of 46 high-elevation lakes receiving 1 to 3.2 kg N/ha/yr in wet deposition, but observed
that diatom communities were dominated by a species associated with moderate N
enrichment even in lakes where NO3 concentrations were at the detection limit
(<1 |ig/L).
New information on relative NO3 inputs from glacial versus snowpack meltwaters
indicate water of glacial origin has higher NO3 , which may influence interpretation of
biological data from high-altitude lakes and streams in some regions of the U.S.,
including the Rocky Mountains (Slemmons et al.. 2015; Slemmons et al.. 2013; Saros et
al.. 2010; Baron et al.. 2009). In the central Rockies and Glacier National Park, fossil
diatom richness in snowpack-fed lakes was found to be higher relative to lakes fed by
both glacial and snowpack meltwaters (Saros et al.. 2010). Sedimentary diatom analysis
of alpine lakes in the Rocky Mountains indicated that increases in A. formosa occurred at
least a century ago in glacially fed lakes whereas shifts in the planktonic diatom
communities occurred after 1970 in snow-fed lakes (Slemmons et al.. 2017).
16.5.5. Rocky Mountain National Park Initiative
The large body of research on the effects of atmospheric deposition on terrestrial and
aquatic ecosystems in ROMO and the relatively low CLs has led the National Park
Service (NPS) to partner with the Colorado Department of Public Health and
Environment and the U.S. EPA to develop a plan in 2005 to decrease atmospheric N
pollution within ROMO (Porter and Johnson. 2007). Formally, this collaboration is
organized as the Rocky Mountain National Park Initiative (RMNPI). The goal of the
collaboration is to understand the impacts of N deposition in ROMO and the sources of N
emissions, and resolve the N deposition issue in ROMO (RMNP Initiative. 2014). The
three agencies developed the Nitrogen Deposition Reduction Plan, which intended to
June 2018
16-138
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
decrease N wet deposition in ROMO from a 2006 baseline of 3.1 kg N/ha/yr to a 2032
target of 1.5 kg N/ha/yr.
The Nitrogen Deposition Reduction Plan aims for a linear decrease in deposition between
2006 and 2032, with interim milestones of 0.3 kg N/ha reductions in annual deposition
every 5 years. In 2012, the 5-year rolling average of wet deposition was 2.9 kg N/ha/yr, a
number that was 0.2 kg N/ha/yr above the interim milestone target, but also reflective of
a 4-year trend toward decreased wet N deposition (RMNP Initiative. 2014). However, in
the most recent status report (Morris. 2016). the 5-year rolling average has increased in
each of the last 2 years and is now 3.3 kg N/ha/yr (Figure 16-42). well above the intended
target. There has been no overall trend in wet N deposition or nitrate deposition since
2007, but the ammonium concentration has increased at four of the five monitoring sites
since 2009. Within the nine-county Denver Metropolitan and Northern Colorado Front
Range region, the two largest estimated sources of ammonia emissions were livestock
(75% of emissions) and fertilizer use [14% (RMNP Initiative. 2015)1. However,
considerable uncertainty remains in ammonia emissions and transport estimates, and the
members of RMNPI have dedicated resources to research ammonia in the region. The
agency partners within the RMNPI are working with the agricultural industry in Colorado
to decrease ammonia emissions by adopting a set of best management practices for
agricultural N (Figure 16-43).
16-139
DRAFT: Do Not Cite or Quote

-------
3.30 kg N/ha/yr*
-2032 Gfidepath
•5-yr Rolling Average Wet N Deposition
2nd Interim Milestone
(2.4 Kg N/ha/yr)
3rd Interim Milestone
{2,1 kg N/ha/yr)
1st Interim Milestone
(2.7 kg N/ha/yr)
4th Interim Milestone
(1.8 kg N/ha/yr)
Resource Management Goal
(1.5 kg N/ha/yr)
Natural Conditions
(0.2 kg N/ha/yr)

8
o
(M
Year
551
a
ha = hectare; kg = kilogram; N = nitrogen; yr = year.
Source: (Morris et al.. 2014).
Figure 16-42 Rocky Mountain National Park Initiative glidepath and current wet
nitrogen deposition at Loch Vale in Rocky Mountain National
Park.
June 2018
16-140
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
RMXP INITIATIVE ACCOMPLISHMENT TIMELINE
•	Established resource
management goal (critical
load)
•	RoMANS field campaign
Evaluate Loch Vale
monitoring data
Critical load interim goal
to be reviewed with
respect to XDRP and CP
Ag = agriculture; APCD = Air Pollution Control District; AQ = air quality; AQCC = Air Quality Control Commission; BMP = best
management practice; CDPHE = Colorado Department of Public Health and Environment; CLA = Colorado Livestock Association;
CP = Contingency Plan; EPA = U.S. Environmental Protection Agency; NADP = National Atmospheric Deposition Program;
NPS = National Park Service; NDRP = Nitrogen Deposition Reduction Plan; RMNP = Rocky Mountain National Park;
RoMANS = Rocky Mountain Atmospheric Nitrogen and Sulfur.
Source: CDPHE (2012). https://www.colorado.gov/pacific/sites/default/files/AP PO ROMO-lnitiative-Accomplishment-Timeline.pdf.
Figure 16-43 Rocky Mountain National Park Initiative accomplishment timeline.
16.5.6. Interactions between Nitrogen Deposition, Precipitation, and
Large-Scale Ecological Disturbances
Precipitation patterns have changed over the past decades in ROMO, with consequences
for biodiversity responses to N deposition. Spring snowfall has diminished over the last
two decades (Clow et al.. 2016). while winter snowfall has increased over the last century
(Concilio et al.. 2016). As a consequence, snowmelt occurs 7-18 days earlier than in
1993 (Clow et al.. 2016). which extends the growing season. In a review of Niwot Ridge
research. Bowman et al. (2014) suggests that N additions tend to have the opposite effect
that high snowpack has on soil properties and processes (i.e., if N deposition accelerates
soil N cycling, higher snowpack slows soil N cycling). Long-term trends in decreasing
snowpack caused by changes in temperature and dust deposition (Clow et al.. 2016). will
intensify the effects of N deposition upon biodiversity.
June 2018
16-141
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
Most forests within ROMO are composed of lodgepole pine or a mixture of Engelmann
spruce and subalpine fir. Historically, these forests have been vulnerable to two natural
mortality agents: bark beetles and wildfire (Ehlc and Baker. 2003). Bark beetles can kill
more than 90% of mature trees in heavily infested areas (Rhoades et al.. 2013). while
wildfires in these types of forests tend to burn at high intensities and kill trees across
large landscapes (Sibold et al.. 2006; Buechling and Baker. 2004; Veblen et al.. 1994;
Romme and Knight. 1981). The large number of trees killed in these disturbances
radically alters many ecological processes, including N cycling.
After tree mortality events, the decrease in biotic demand for N results in higher levels of
inorganic N in the soil and increased concentrations of N in the leaves of surviving or
recolonizing plants (Diikstra and Adams. 2015; Mikkelson et al.. 2013; Rhoades et al..
2013). These changes in N cycling are similar to those caused by N deposition (Fcnn et
al.. 1998) and those in experiments in and near ROMO simulating N deposition (Rueth et
al.. 2003). Although these disturbances increase N availability, beetle epidemics or
wildfires in Rocky Mountain lodgepole pine and spruce-fir forests have not historically
caused large increases in soil nitrate leaching or high concentrations of N in stream water
(Dunncttc et al.. 2014; Morris et al.. 2013; Rhoades et al.. 2013). processes associated
with excess N availability and the state of "N saturation" that occurs in ecosystems
receiving large amounts of N deposition (Fcnn et al.. 2003a). However, there is concern
that N deposition along the Colorado Front Range will enhance N availability to the point
where disturbances will lead to large N leaching losses (Rhoades et al.. 2013). This
means that there is potential for N deposition to interact with bark beetles and wildfire to
increase stream and lake water N content and degrade ecosystem function.
Nitrogen deposition can increase insect attack on plants (Throop and Lerdau. 2004) and
there is some evidence that bark beetle activity and the bark beetle-caused tree mortality
can be increased by N deposition (Jones et al.. 2004). Unlike in other parts of the western
U.S. where N deposition has allowed fires to occur more frequently by increasing fuels
through greater plant growth (Fenn et al.. 2003a). there is unlikely to be a large effect of
N deposition on fire occurrence in ROMO. Fires in ROMO occur mostly during extreme
weather conditions and are not as directly influenced by the amount of fuel available
(Sibold et al.. 2006). Notably, warm climate conditions increase fire frequency in
lodgepole pine and spruce-fir forests in the Rocky Mountains (Westerling et al.. 2006)
and support bark beetle outbreaks in these forests (Bentz et al.. 2010; Hebertson and
Jenkins. 2008).
16-142
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
16.6.
Southern California
16.6.1. Background
This case study focuses on the impact of N and S deposition on aquatic and terrestrial
ecosystems in southern California. Two areas are considered that have abundant
information on ecological effects: the Sierra Nevada and southern California arid land
ecosystems. Here we focus on the national parks located in both of these ecosystems, the
Sequoia and Kings Canyon national parks (SEKI), located in the southern Sierra Nevada
Mountains in southern California, and Joshua Tree National Park (JOTR), located in the
desert to the southeast (Figure 16-44). SEKI and JOTR were selected to emphasize
Class I areas in the Sierra Nevada (SEKI) and southern California arid land ecosystems
(JOTR). In addition, research from surrounding ecoregions relevant to these ecosystems
is included because more work has been conducted within the surrounding areas than
within the park boundaries.
The majority of the information provided in this case study represents the Sierra Nevada
Mountains/SEKI. Limited research conducted in JOTR was found in the peer-reviewed
literature. Yosemite National Park (YOSE) to the north shares many commonalities with
SEKI regarding air pollution exposure, sensitivities, and impacts. Thus, research
conducted in YOSE is relevant and included here. Background information is provided
below, including a general description of the region, protected status designation
(e.g., Class I, Wilderness Area), and regional land use. This case study includes
atmospheric deposition to the parks and surrounding areas, dose-response relationships,
and critical loads.
June 2018
16-143
DRAFT: Do Not Cite or Quote

-------
ALPINE
CHAPARRAL
COASTAL SAOF. SCRUB
DESERT
¦ GRASSLAND
MIXED CONIFER
¦	MIXED liARDWOOD
¦	OAK WOODLAND
RIPARIAN
SAGEBRUSH STEPPE
m WATER
WETLAND
Mta
¦hi ii' uriLti i«n
Cm tm ('inri<— hniim
Ut -	W
Notes: Land cover presented in this figure represents potential natural vegetation before urbanization and modern agricultural
development.
Source: Reclamation et al. (1996V from Fenn et al. (2010V Adapted to show case study locations.
Figure 16-44 Map of the distribution of vegetation types and land cover in
California.
June 2018	16-144	DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
16.6.1.1.
Description of Case Study Region
Sequoia and Kings Canyon national parks are located in the Mediterranean California
ecoregion (Level I Omernick classification) and are part of the Sierra Nevada Mountains.
These parks are known for exceptionally large trees, high mountain peaks and ridges, and
deep canyons and include the highest point in the contiguous U.S. (CONUS), Mount
Whitney at 14,494 ft (4,418 m) above sea level. There are more than 200 marble caverns,
many with endemic cave invertebrates. Sequoia National Park was originally created in
1890 to protect the giant sequoia trees (Sequoiadendron giganteum) from logging. It was
the second national park and the first formed to protect a particular living organism. In
1940, Congress created the contiguous Kings Canyon National Park to include the
glacially formed Kings Canyon. These parks have increased in size to encompass
1,353 mi2 (3,504 km2); 97% is designated and managed as wilderness.
Ecosystems in SEKI include chaparral vegetation at low elevation, extending up to alpine
areas in the high mountains. Mycorrhizae, lichens, and herbaceous species communities
are susceptible to declines in species richness/biodiversity due to the effects of N
deposition. Likewise, high-elevation lakes and streams as well as low-order streams are
susceptible to eutrophication and acidification driven by N deposition. Five rivers begin
in SEKI, and past and present glaciers contributed to more than 3,000 lakes (many at high
elevation) and 2,000 miles of streams, generally of low Strahler order (Boiano ct al..
2005). Landscape variation contributes to diverse habitats that contain various
assemblages of terrestrial, aquatic, and subterranean biota. As of 2005, of the more than
1,500 taxa of vascular plants in SEKI, 138 were deemed special status; 24 were listed by
the California Native Plant Society as rare, threatened, or endangered. Fifty-one taxa of
the parks' 312 terrestrial and aquatic vertebrate species had special status designations.
Five extant native vertebrate species were federally listed in 2005: Sierra Nevada bighorn
sheep (Ovis canadensis sierrae), bald eagle (Haliaeetus leucocephalus), Little Kern
golden trout (Oncorhynchus mykiss whitei), mountain yellow-legged frog (Rana
muscosa), and Yosemite toad (Anaxyrus canorus). As of 2005, the state listed the
peregrine falcon (Falco peregrinus), great gray owl (Strix nebulosa), and willow
flycatcher (Empidonax traillii) as endangered and the red fox (Vulpes vulpes), wolverine
(Gulo gulo), and Swainson's hawk (Buteo swainsoni) as threatened. Forty extant native
vertebrate species were listed as sensitive, including: white-tailed jackrabbit (Lepus
townsendii), fisher (Martespennanti), ring-tailed cat (Bassariscus astutus), Townsend's
big-eared bat (Corynorhinus townsendii), western mastiff bats (Eumops perotis), merlin
(Falco columbarius), northern harrier (Circus cyaneus), California spotted owl (Strix
occidentalis), California legless lizard (Anniellapulchra), western pond turtle (Actinemys
marmorata), Mount Lyell salamander (Hydromantes platycephalus), San Joaquin roach
June 2018
16-145
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
fish (Lavinia symmetricus), and Kern River rainbow trout [Oncorhynchus mykiss gilberti;
(Boiano et al.. 2005)1.
Joshua Tree National Park is in the North American Desert ecoregion (Level I ecoregion
by Omernick classification). The park was initially dedicated as a U.S. National
Monument in 1936 and later declared a national park in 1994 when the U.S. Congress
passed the California Desert Protection Act (Public Law 103-433). The park covers
1,235 mi2 (3,200 km2), more than half designated as wilderness. It straddles the San
Bernardino County/Riverside County border, including parts of two deserts: the higher,
wetter Mojave Desert and the lower, drier Colorado Desert. The Little San Bernardino
Mountains cross the southwestern edge of the park. JOTR is covered primarily by
semiarid and arid vegetation types, including Joshua tree (Yucca brevifolia) woodlands.
While rainfall is less than 10 cm/year, JOTR's groundwater supplies 200 surface water
sources (e.g., springs, oases, ephemeral streams, or washes).
Joshua Tree National Park was preserved to protect the natural resources found at the
junction of three California ecosystems. The Colorado Desert is a western extension of
the Sonoran Desert. It occupies the southern and eastern portions of the park, and is
characterized by ocotillo (Fouquieria splendens) shrubs and cholla (Opuntia spp.) cacti.
The Mojave Desert crosses the northern portions of the park and contains Joshua tree
woodlands. The third ecosystem type in JOTR is a community of California juniper
(Juniperus californica) and pinyon pine (Pinus spp.) located in the westernmost portion
of the park, above 1,220 m elevation, in the Little San Bernardino Mountains.
There are several threatened and endangered species found within JOTR. Plant species
include the triple-ribbed milkvetch (Astragalus tricarinatus) and Coachella Valley
milkvetch (A. lentiginosus var. coachellae), which are federally endangered; as well as
the Parish's daisy (Erigeronparishii), which is federally threatened. The desert tortoise
(Gopherus agassizii), California's state reptile, is on both California's and the federal
threatened lists. Twenty-six other species are listed as "special concern." JOTR was
originally inhabited by the Serrano, the Chemehuevi (sometimes called the Southern
Paiutes), and the Cahuilla people, and the park contains over 500 archaeological sites.
(https://www.nps.gOv/i otr/index.htm').
16.6.1.2. Class I Areas
The 1964 Wilderness Act and the NPS Organic Act are both used as tools to protect
SEKI and JOTR from air pollution impacts. SEKI and JOTR (as well as nearby YOSE)
are also Prevention of Significant Deterioration (PSD) Class I areas. They receive the
highest level of protection under the Clean Air Act (CAA) against air pollution
June 2018
16-146
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
degradation. The CAA (42 USC 7470) authorized designation of Class I areas to protect
air quality in national parks (over 6,000 ac [2,428 hectares]) and national wilderness
areas (over 5,000 ac [2,023 hectares]) in an effort to preserve pristine atmospheric
conditions and air quality-related values (AQRVs). Class I areas are subject to the PSD
regulations under the CAA. PSD preconstruction permits are required for new and
modified existing air pollution sources. Air regulatory agencies are required to notify
federal land managers (FLMs) of any PSD permit applications for facilities within
100 km of a Class I area.27 The FLMs are authorized to review and comment on PSD
Class I permit applications with the permitting agency.
16.6.1.3. Regional Land Use and Land Cover
Land cover in SEKI is largely forested and mountainous terrain in and near the Sierra
Nevada Mountains. The westernmost portions of the park are covered by mixed conifer
and chaparral vegetation. Further to the west is a zone of intensive agriculture
(Table 16-29).
Table 16-29 Land coverages of Sequoia, Kings Canyons, and Joshua Tree
national parks.
Land Cover Category
Sequoia National Park
(km2)
Joshua Tree National Park
(km2)
Kings Canyon
National Park
(km2)
Developed
5
7
4
Barren land and exposed
rock
416
207
690
Deciduous forest
12
0
2
Evergreen forest
736
5
519
Mixed forest
8
0
1
Shrub/scrub
390
2,936
516
Grassland/herbaceous
68
54
100
Wetland
2
0
2
km = kilometer.
27 http://webcam.srs.fs.fed.us/psd.
June 2018
16-147
DRAFT: Do Not Cite or Quote

-------
Land cover in JOTR is largely desert and semiarid land (Table 16-29. Figure 16-44).
Nearby human population centers (Figure 16-45) include the Coachella Valley
(-350,000 people, including the cities of Palm Springs and Indio), as well as the towns of
Twentynine Palms (25,600 people) and Joshua Tree (7,400).
Hi
V
\
w •
r
\
i
I
r \
¦ L "I.	r)
J Vosennita

.Sierra National
Forest
Sequoia
National Park
Kind's Canyon
National Park
NEVADA
CALIFORNIA
Los Padres
National Forest
"55s

Santa Monica
Mountains
Southern California
Case Study Region
Case Study Areas
Native American
Reservations
USA City Populations
•	1,000,000 plus
•	500,000 - 999,999
250,000 - 499,999
•	100,000 - 249,999
20 40 60
CP
)	y\
Mojave National 4^;
Preserve
v\
Figure 16-45 Southern California case study region showing locations of
human population centers.
June 2018
16-148
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
16.6.2. Deposition
Characteristics of N and S deposition are shown in Figure 16-46 through Figure 16-48 for
JOTR and Figure 16-49 through Figure 16-51 for SEKI. Data shown in the figures were
obtained from the hybrid modeling/data fusion product, TDEP
(http://nadp.sws .uiuc.edu/committees/tdcp/tdcpmaps/). and described earlier in
Appendix 2. However, the time series of wet deposition is taken directly from data on the
NADP/NTN website. This was done to track changes in deposition since the passage of
the Clean Air Act Amendment because the CMAQ simulations involved in TDEP extend
back only to 2000. See Appendix 2.4. Appendix 25_ and Appendix 2JS for more
information on deposition in the U.S. Other maps showing the contributions of individual
species to dry and/or wet deposition are given in Appendix 2.7.
Figure 16-46 indicates that the general pattern of deposition in JOTR of N and S is
broadly similar, tending to be higher near the NADP site located in the NW corner of the
study area and decreasing inland. As expected, N deposition is highest in and surrounding
Los Angeles. The park is often subject to transport of emissions from the Los Angeles
Basin. However, as can be seen from Figure 16-46B. this influence decreases
substantially towards the southeast. In addition, the area surrounding JOTR shows a high
degree of regional heterogeneity. Deposition of S is consistent with the rest of the
western U.S., but is considerably lower than in the eastern U.S.
As shown in Figure 16-47A. deposition of N is estimated to be mostly in oxidized forms
in JOTR. Although there are areas principally to the south or to the west where N is
deposited mostly in reduced forms. In Figure 16-48. wet deposition of NO3 , NH4 .
SO42 . and H+, apart from being lower than at many other sites, has not shown consistent
trends over the past 25 years. Comparison of Figure 16-46 and Figure 16-47 shows that
dry deposition dominates over wet deposition of N and S in the JOTR study area.
In SEKI, Figure 16-49 shows a 3-year average total deposition of N and S for
2011-2013; Figure 16-50 shows the partitioning between oxidized and total N;
Figure 16-51 shows the 25-year time series for wet deposition of NO3 , NH4 . SO42 . and
H+ obtained at the NADP/NTN monitoring sites near SEKI.
June 2018
16-149
DRAFT: Do Not Cite or Quote

-------
~ LO* Ar»gal*v CA
O Mon'tof CA ( 6?
) Monitor location*
Jo*ho» Tr« Study *'«
S Dwmw l«e-S4t*)
•> S*» V* «>
5^v»w'
*
V
Lo* Ang#l#v CA

O Monitor CA ¦ S7

# Monitor Location*

Jothu* Tree Study *¦'»'




¦ «¦

0
N OqKMDWn (Ifl H'M)
'VV^VVV*
* * % \v so
CONUS = contiguous U.S.; N = nitrogen; S = sulfur; ha = hectare; kg = kilogram.
Surrounding areas in California and Nevada and inserts showing the whole CONUS are shown to place the depositional environment for both portions of the study area in context.
Other maps showing the contributions of individual species to dry and/or wet deposition are given in Appendix 2.7.
Figure 16-46
Patterns and temporal trends of nitrogen and sulfur deposition in Joshua Tree National Park and
surrounding region in California. A and B show the 3-year average total deposition of nitrogen
and sulfur for 2011 -2013
June 2018
16-150
DRAFT: Do Not Cite or Quote

-------
CONUS = contiguous U.S.; N = Nitrogen, S = Sulfur.
Surrounding areas in California and Nevada and inserts showing the whole CONUS are shown to place the depositional environment for both portions of the study area in context.
Other maps showing the contributions of individual species to dry and/or wet deposition are given in Appendix 2,7,
Figure 16-47 Patterns and temporal trends of nitrogen and sulfur deposition in Joshua Tree National Park and
surrounding region in California. A shows the partitioning between oxidized and reduced
nitrogen; B and C show the 3-year average total percentage of wet deposition of nitrogen and
sulfur for 2011-2013.
June 2018
16-151
DRAFT: Do Not Cite or Quote

-------

100

90


L
>
80
T-l

CD
70
JO

~o
60
E


50
c

o
•
40
+-•

'u~>
o
30
Q.

0)
20
a


10

0
Annual Wet Deposition and 3-Year Moving Average at
Site CA67: 1990-2014
•	NIV
•	NOr
•	so,1-
H+ Lab
C1
/ \
/ \
/
I
* /
7 /N




Ł
~'-i r-^
1 • 		
•\
1999
2001
2003
2005
2007
Year
2009
2011
. _ _
• ~'
2013
2015
H+ = hydrogen ion; ha = hectare; mol = mole; NH4+ = ammonium; N03 = nitrate; S042 = sulfate; yr = year.
Surrounding areas in California and Nevada and inserts showing the whole CONUS are shown to place the depositional environment for both portions of the study area in context.
Other maps showing the contributions of individual species to dry and/or wet deposition are given in Appendix 2.7.
Figure 16-48
Patterns and temporal trends of nitrogen and sulfur deposition in Joshua Tree National Park and
surrounding region in California. The 25-year time series for wet deposition of nitrate,
ammonium, sulfate, and hydrogen ion obtained from the National Atmospheric Deposition
Program/National Trends Network monitoring site CA 67.
June 2018
16-152
DRAFT: Do Not Cite or Quote

-------
CONUS = contiguous U.S.; H+ = hydrogen ion; ha = hectare; mol = moie; NH/ = ammonium; NO32" = nitrate; S042 = sulfate; yr = year.
Surrounding areas in California and Nevada and inserts showing the whole CONUS are shown to place the depositional environment for both portions of the study area in context.
Other maps showing the contributions of individual species to dry and/or wet deposition are given in Appendix 11.
Figure 16-49 Patterns and temporal trends of nitrogen and sulfur deposition of Sequoia and Kings Canyons
national parks and surrounding region in California. A and B show the 3-year average total
deposition of nitrogen and sulfur for 2011 -2013.
June 2018
16-153
DRAFT: Do Not Cite or Quote

-------
CONUS = contiguous U.S.; H+ = hydrogen ion; ha = hectare; mol = mole; NH4+ = ammonium; N032" = nitrate; S042 = sulfate; yr = year.
Surrounding areas in California and Nevada and inserts showing the whole CONUS are shown to place the depositional environment for both portions of the study area in context.
Other maps showing the contributions of individual species to dry and/or wet deposition are given in Appendix 2,7,
Figure 16-50 Patterns and temporal trends of nitrogen and sulfur deposition of Sequoia and Kings Canyons
national parks and surrounding region in California. A shows the partitioning between oxidized
and reduced nitrogen, indicated as the fraction of total nitrogen which is oxidized; B and C show
the 3-year average total percentage of wet deposition of nitrogen and sulfur for 2011-2013.
June 2018
16-154
DRAFT: Do Not Cite or Quote

-------
Annual Wet Deposition and 3-Year Moving Average at A
ivo
too
no
2
^ MO
8 ivo
1*00
a v)
Site CA75: 1990 - 2014
•*»C
•	NO,
•	«>,*
• H* i Ji
N
• %
S-V
MM
1 • ir, • tv^'v
iw im /ooo M4 ma Mil
Year
V
*>16
Annual Wtt Deposition and 3-Year Moving Average at B
TOO
1M
>. 1M
\ 140
XL
-» uo
JOO
•5 10
S to
V 40
JO
0
1H
Sit# CA99: 1990 2014
•	HO.
•	SO,^
•	N* 1*6
/% • / \
/ # \
•	/--'If	%V ' • \ #	^
#	• *Y / # • k	\#/*
V..... • . ,\vw.
! •./ . . V..	*v	 '•

• • . • v«" . • *u
l*W
MOO	M04
Yetr

MU
CONUS = contiguous U.S.; H+ = hydrogen ion; ha = hectare; moi = moie; NH4+ = ammonium; N032" = nitrate; S042 = sulfate; yr = year.
Surrounding areas in California and Nevada and inserts showing the whole CONUS are shown to place the depositional environment for both portions of the study area in context.
Other maps showing the contributions of individual species to dry and/or wet deposition are given in Appendix 2.7.
Figure 16-51 Patterns and temporal trends of nitrogen and sulfur deposition of Sequoia and Kings Canyons
national parks and surrounding region in California. A and B show the 25-year time series for wet
deposition of nitrate, ammonium, sulfate, and hydrogen obtained from the National Atmospheric
Deposition Program/National Trends Network monitoring sites CA99 and CA75.
June 2018
16-155
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
Deposition of N is much higher in SEKI than JOTR. There is also much greater spatial
variability in N deposition in SEKI than JOTR. On the other hand, S deposition is similar
in magnitude between the two sites; S is also characterized by relatively low spatial
variability at the two sites. In contrast to JOTR, deposition of N is estimated to be mostly
in reduced forms in SEKI (see Figure 16-50). This is not surprising, given the proximity
of SEKI to the San Joaquin Valley. There is considerable variability in the percentage of
N deposition as either in oxidized or reduced forms. Wet deposition of NO3 , NH4+,
SOr . and H+ is larger over SEKI than over JOTR, but still lower than at some NADP
sites in the eastern U.S. Similar to JOTR, dry deposition dominates over wet deposition
of both N and S in SEKI.
A recent modeling study of Class I areas, including SEKI and JOTR, used a GEOS-Chem
adjoint model to identify the geographic sources of reactive N deposition, as well as the
emission sector sources of reactive N deposition within the parks (Lee et al.. 2016). 90%
of emissions of Nr that affect SEKI originate within 400 km of the park, with livestock
ammonia as the major emission source (>50%) of N deposition within SEKI (see
Figure 16-52). At JOTR, 90% of N emissions originate within 600 km of the park, and
mobile sources of NOx are the major emission source (>60%) of N deposition to JOTR
(see Figure 16-52.
16-156
DRAFT: Do Not Cite or Quote

-------
Geographic footprint of Nr deposition
0.005 0.01 0.02 0.03 0.04 0.06 0.08 0.1
[kg N/ha/yr]
Emission sectors
NOx(surface inventory)
| NOx(electric generating units)
NOx (non-electric generating
industrial stacks)
J NOx (aircraft)
NH3(livestock)
H NH3 (fertilizer)
^ NH3(natural)
NOx(lightning)
^ NOx (soil)
Figure 18-52 Annual-averaged monthly footprint of reactive N deposition in
Joshua Tree National Park (3.2 kg N/ha/yr) and Sequoia National
Park (5.7 kg N/ha/yr). Also shown for each park is a pie chart of
fractional contribution from emission sectors, as estimated by
GEOS-Chem adjoint model.
16.6.3. Critical Loads and Other Dose-Response Relationships
16.6.3.1. Terrestrial
1	The communities of terrestrial species with well-documented effects caused by N
2	deposition include lichens, mycorrhizae, grasses, and trees. While well studied
3	ecosystems include mixed conifer forest, alpine/subalpine, and chaparral/semiarid
4	shrubland.
5	Data on the effects of N deposition on mixed conifer forest ecosystems in the
6	Mediterranean California ecoregion (Omernick Level I), including the Sierra Nevada
7	region, are well summarized by Fenn et al. (201 lb). A more recent overview of empirical
8	and modeled CLs for mixed conifer forests was provided by Fenn et al. (2015). Nitrate
June 2018
16-157
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
leaching from soil in combination with high evapotranspiration rates lead to high nitrate
concentrations in surface waters in this area.
In general, the alpine and subalpine plant communities, such as those found in SEKI and
YOSE, are sensitive to eutrophication, an observation based largely on research
conducted in alpine ecosystems in the Rocky Mountains (Bowman et al.. 2006). Alpine
and subalpine communities are high-elevation plant communities that have developed
under conditions of low nutrient supply, in part because soil-forming processes tend to be
poorly developed at high elevations. This contributes to N sensitivity. Consequently,
changes in alpine plant productivity and species composition have been reported in
response to increased N inputs (Bowman et al.. 2006; Vitousek et al.. 1997).
In chaparral and other semiarid shrublands in southern California, nitrogen enrichment
can contribute to an increase in net primary productivity; however, water availability is
often limiting to growth (Vourlitis. 2012). Arid ecosystems represent an extreme in terms
of water availability and water is the principal limiting resource. Therefore, long-term
(many years) experiments and the potential for interactions among water, fire, and
nutrient supply in governing plant responses to N additions should be considered. Based
on N addition experiments, Vourlitis and Pasquini (2009) suggested that dry-season
addition of N significantly changes the community composition of coastal sage scrub
(CSS) vegetation, but not chaparral. The impacts of such changes on the plant community
may magnify over time. It has been proposed that increases in both fire frequency and N
deposition over the last several decades in southern California may have promoted the
conversion of CSS land to non-native grasslands (Talluto and Suding. 2008). These same
effects may also occur in other arid ecosystems including JOTR. Both empirical and
modeling studies suggest that loss of biodiversity in response to N input and climate
change can be greater than the effects of either stressor alone (Porter et al.. 2013).
Empirical Studies
A number of empirical studies (N gradient and N experimental additions) have been
conducted to identify tipping points and to define areas of CL exceedance. Key studies
conducted in the southern California case study region are listed in Table 16-30.
Fenn et al. (2010) reported empirical CL exceedance maps for seven major vegetation
community types in California (Figure 16-53). Thirty-five percent of the land area
covered by these vegetation types (100,000 km2) was estimated to be in exceedance of
the nutrient-N CLs (3-8 kg N/ha/yr) that were estimated for mixed conifer forests,
chaparral, and oak woodlands. Nearly half of the area covered by CSS (54%) and
grasslands (44%) was judged by the authors to be in exceedance of the CLs for protecting
16-158
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
the ecosystem against changes associated with invasive grasses. Substantial chaparral
(53%) and oak woodland (41%) cover was estimated to be in exceedance for impacts on
epiphytic lichens. About 30% of the desert area investigated was in exceedance, based on
the presence of invasive grasses and increased fire risk. An estimated 30% of the mixed
conifer forest, based on lichens, was in exceedance of the CL. Forested and chaparral
communities were less sensitive. Only 3-15% of the forested and chaparral areas were
estimated to be in exceedance of the CL for NOs leaching. By combining the exceedance
areas of the seven vegetation types, Fenn et al. (2010) estimated that 25% of the land area
of the vegetation types included in the study were in exceedance for protecting against
nutrient enrichment.
Pardo etal. (2011c) conducted a synthesis study that evaluated published data to identify
the empirical CLs for N in Level I ecoregions across the U.S. The authors estimated CL
values as low as 3 kg N/ha/yr to protect sensitive resources in the southern California
case study area. Table 16-30 lists the CLs identified by Pardo etal. (2011c) for North
American Desert and Mediterranean California ecoregions (Omernick Level I) and new
studies published after Pardo et al. (2011c). If deposition exceeds the CL, the risk of
harmful effects increases.
16-159
DRAFT: Do Not Cite or Quote

-------
< umfwrMir < ritk^l I i uhI:
IliKb-fnd Ihmkvkl I irmlvncf
CL = critical load; ha = hectare; kg = kilogram; N = nitrogen; yr = year.
When CLs based on two different N responders were used within a given vegetation type (e.g., lichen effects and nitrate leaching),
only the more sensitive responder was used (lowest CL). The higher value of two CLs, when two values were available, was used in
the case of coastal sage scrub, mixed conifer forest (lichen community effects), desert scrub and pinyon-juniper, and the lichen
effects CL was used for chaparral.
Source: Fenn et al. (2010). Adapted to show case study locations.
Figure 16-53 Composite critical load exceedance maps for all seven vegetation
types included in the study of Fenn et al. (2010) showing the
combined exceedance areas and the level of exceedance (kg
N/ha/yr).
June 2018
16-160
DRAFT: Do Not Cite or Quote

-------
Table 16-30 Summary of recent empirical dose-response and critical load studies focused on the southern
California case study area and published since Pardo et al. (2011c).
Study Terrestrial	Aquatic	Location	Focus	CL/Exceedance
Mediterranean California Ecoregion Level I (Omernick)
3.1-39 kg N/ha/yr
The lowest critical load is 3.1
based on lichen tissue chemistry
above the clean site threshold.
Nitrate leaching and fine root
biomass critical load is 17, soil
acidification is 26 kg N/ha/yr,
and susceptibility to beetle
infestation is 39 kg N/ha/yr
3.1-14 kg N/ha/yr
3.1 kg N/ha/yr is a modeled
value for lichens, while
10 kg N/ha/yr is the value for
nitrate leaching
Pardo et al.
(2011a)
•
Mediterranean California
Ecoregion Level I (Omernick);
coastal sage scrub ecosystem
Exotic invasive grass cover,
native forb richness, arbuscular
mycorrhizal richness
7.8-10 kg N/ha/yr
Pardo et al.
(2011a)
•
Mediterranean California
Ecoregion Level I (Omernick);
serpentine grassland ecosystem
Annual grass invasion replacing
native herbs
6 kg N/ha/yr
Clark et al.
(2013)
•
Mediterranean California and
North American Desert
Loss of herbaceous plant
species
Species loss 1-30%
Ellis et al.
(2013)
•
SEKI and YOSE
Protection of lichens
Estimated CL 2.5 to
7.1 kg N/ha/yr
Pardo et al. •	Mediterranean California	Lichen chemistry and community
(2011a)	Ecoregion Level I (Omernick); changes, nitrate leaching, soil
mixed conifer forest ecosystem acidification, reduced fine root
biomass
Pardo et al. •	Mediterranean California	Nitrate leaching and changes in
(2011a)	Ecoregion Level I (Omernick); the lichen community
chapparal ecosystem
June 2018
16-161
DRAFT: Do Not Cite or Quote

-------
Table 16-30 (Continued): Summary of recent empirical dose response and critical load studies focused on the
southern California case study area and published since {Pardo, 2011,
2804014@@author-year}
Study
Terrestrial Aquatic
Location
Focus
CL/Exceedance
Cox et al.
(2014)
•
Mediterranean California
Ecoregion Level I (Omernick);
coastal sage scrub ecosystem
Conversion from coastal sage
scrub to annual grassland
Estimated CL 11 kg N/ha/yrto
conserve coastal sage shrub
vegetation
Bvtnerowicz
et al. (2015)
•
Mediterranean California
Ecoregion Level I (Omernick);
chapparal ecosystem
Lichens in chaparral, San
Bernardino Mountains
Exceedances of
CL = 5.5 kg N/ha/yr for lichens
occurred mostly in western and
northern portions of the study
area
Bvtnerowicz
et al. (2015)
•
Mediterranean California
Ecoregion Level I (Omernick);
mixed conifer forest ecosystem
Mixed conifer forest
Exceedances of
CL = 3.1 kg N/ha/yr for mixed
conifer forest covered much of
San Bernardino Mountains
Allen et al.
(2016)
•
Mediterranean California
Ecoregion Level I (Omernick);
coastal sage scrub ecosystem
Rapid decline in mycorrhizal
biodiversity
CL = 11 kg N/ha/yr
North American Desert Ecoregion Level 1 (Omernick)
Pardo et al.
(2011a)
•
North American Desert
Ecoregion Level I (Omernick)
Protect herbaceous plants and
shrubs
CL 3 to 8.4 kg N/ha/yr
Simkin et
al. (2016)
•
North American Desert
Ecoregion Level I (Omernick)
Grass and forb decreasing
species richness
CL open canopy: 8.3-9.9
(mean = 9.2, n = 240)
CL closed canopy: 13.5-17.0
(mean = 16.5, n = 32)
Clark et al.
(2013)
•
Mediterranean California and
North American Desert
Loss of herbaceous plant
species
Species loss 1-30%
CL = critical load; ha = hectare; kg = kilogram; N = nitrogen; SEKI
= Sequoia and Kings Canyons National Parks; YOSE = Yosemite National Park; yr = year.
June 2018
16-162
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
Data compiled by Pardo etal. (2011c) suggested that ambient N deposition is higher than
the lower limit of the expected CL to protect against nutrient enrichment effects in some
of the national parks in the Sierra Nevada, mainly SEKI, and to a lesser extent, YOSE.
Potential CL exceedances were reported for the protection of mycorrhizal fungi, lichens,
herbaceous plants, and forest vegetation, and also to restrict NOs leaching in drainage
waters. Fenn et al. (2010) estimated low CL values (3-8 kg N/ha/yr) for mixed conifer
forests, chaparral, and oak woodlands. These low critical load estimates were driven by
the presence of highly N sensitive species (lichens) and N poor vegetation (annual
grasslands and desert scrub plant communities). Fenn et al. (2010) concluded that N
deposition at or above these critical load values might cause vegetation community
changes because the increased availability of N tends to favor invasion by non-native
grasses into arid land ecosystems.
Clark et al. (2013) estimated the loss of herbaceous plant species caused by atmospheric
N deposition based on a national CL database. Using the lower and more conservative
estimates of CLs, the ambient N deposition was judged to be in exceedance of nutrient N
CLs over extensive portions of the Mediterranean California and North American Desert
ecoregions. Estimated plant species losses ranged from less than 1 to 30%, but variability
was high and uncertainty in these estimates was substantial.
Modeling Studies
Dynamic and steady-state models have been used to estimate critical or target loads in the
southern California case study region. Key studies, highlighted in Table 16-31. have
focused on CLs of nutrient N and acidity.
16-163
DRAFT: Do Not Cite or Quote

-------
Table 16-31 Terrestrial critical and target load and exceedance modeling studies
in southern California.
Study
Location
Model
Focus
Mixed Conifer
Hurteau et al. (2009)
Teakettle
Experimental Forest
and YOSE
STELLA®7
Modeled how changing precipitation and N deposition
levels affect shrub and herb biomass production. Herb
cover growth rate greater at 12 kg N/ha/yrthan
24 kg N/ha/yr. Precipitation was more important driver for
shrub cover.
Fenn et al. (2015)
Sierra Nevada and
San Bernardino Mts.
DayCent
Modeled CLto protect against NO3" leaching; CL
ranged from 17 to 30 kg N/ha/yr.
Semiarid
Li et al. (2006)
SEKI
DayCent
Simulated N deposition of 7.7 kg N/ha/yr; increased
loss of N to NO3" export and gaseous NO.
Rao et al. (2010)
JOTR
DayCent
Effects on fire risk in creosote bush and pinyon-juniper
vegetation types. CL = 32 kg N/ha/yr (creosote bush)
and 39 kg N/ha/yr (pinyon-juniper). Wet areas having
low soil clay (6-14%) may have CL as low as
1.5 kg N/ha/yr.
Rao et al. (2010)
JOTR
DayCent
Modeled CL was determined as the N deposition load
at which fire risk began to increase exponentially above
the background fire risk. The CL for creosote bush
scrub was 2.1 kg N/ha/yr and for pinyon-juniper
woodland was 3.6 kg N/ha/yr.
CL = critical load; ha = hectare; JOTR = Joshua Tree National Park; kg = kilogram; Mts. = mountains; N = nitrogen; NO = nitric
oxide; N03" = nitrate; SEKI = Sequoia and Kings Canyons National Parks; YOSE = Yosemite National Park; yr = year.
16.6.3.1.2.1. Mixed Conifer
1	Hurteau et al. (2009) developed and tested a model to determine how changes in
2	precipitation and N deposition might affect shrub and herb biomass in mixed conifer
3	forests in Teakettle Experimental Forest (located between SEKI and YOSE) and YOSE.
4	They estimated the prescribed fire intervals needed to counteract high fuel loads.
5	Increased understory (especially grass) biomass enhances fuel connectivity and fosters
6	larger and more severe fires. Under a model scenario that specified higher interannual
7	variability in precipitation and increased N deposition, model results suggested that
8	implementing fire treatments at an interval approximately equivalent to the historical
June 2018
16-164
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
range (15-30 years) would likely be needed to maintain understory vegetation fuel loads
at levels comparable to the control.
Based on 28 streams draining California mixed conifer forests studied by Fenn and Poth
(1999). Fenn et al. (2015) estimated that peak NOs concentrations in runoff were usually
less than 14.3 |icq/L: therefore, this level of NO3 leaching was identified as the likely
critical threshold value to identify watersheds that exhibit signs of N saturation. Fenn et
al. (2015) specified the empirical CL for protecting against NO3 leaching in California
mixed conifer forests equal to the N deposition level at which this identified critical
surface water NO3 concentration is exceeded (Fenn et al.. 2015; Fenn et al.. 2008). The
CL was derived using linear regression of stream water NO3 concentrations during the
winter high flow period and annual throughfall N deposition at 11 locations in the
southern Sierra Nevada and San Bernardino mountains. The calculated CL to protect
against NO3 leaching at the sites (6-71 kg N/ha/yr deposition) was 17 kg N/ha/yr (95%
CI 15-19 kg N/ha/yr).
16.6.3.1.2.2. Semiarid Shrublands
Li et al. (2006) used the DayCent model to quantify water, C, and N fluxes for a
chaparral ecosystem in SEKI that received about 7.7 kg N/ha/yr of atmospheric N
deposition. Simulated NO3 export and gaseous N emissions (mostly as nitric oxide
[NO]) generally agreed with observations. The model projections suggested that
increased N deposition would likely increase the flux of N from terrestrial ecosystems as
N03 in streams and gaseous NO. The authors concluded that the representations within
the DayCent model of N mineralization, runoff, and NO3 export in chaparral or similar
semiarid vegetation needed improvement.
Rao et al. (2010) used DayCent to estimate CLs of N deposition in JOTR, with focus on
the effects of N input on fire risk in creosote bush (Larrea tridentata) and pinyon-juniper
vegetation communities. Fire risk was expressed as the probability that annual biomass
production would exceed the risk threshold of 1,000 kg/ha of biomass available for
burning. Critical load was calculated as the N deposition at the point along the deposition
gradient where modeled fire risk began to increase exponentially. Mean estimated CLs
across multiple soil types, receiving precipitation of less than 21 cm/year, were 3.2 and
3.9 kg N/ha/yr for creosote bush and pinyon-juniper plant communities, respectively.
Critical loads decreased (more nutrient-sensitive) with decreasing soil clay content and
increasing precipitation. The wettest areas that had low clay content (6 to 14%) had
estimated low CLs, as low as 1.5 kg N/ha/yr (Rao et al.. 2010). These values fall at the
lower end of the CL range identified by Pardo et al. (2011c) and Pardo et al. (2011a) for
herbaceous vegetation in North American Deserts. Fire risks in the two vegetation types
June 2018
16-165
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
were highest under moderately high N deposition (9.3 and 8.7 kg N/ha/yr, respectively);
above those deposition levels, modeled fire risk was influenced more by precipitation
amount than by N supply.
16.6.3.1.2.3. Arid/Semiarid
DayCent modeling results suggested a CL less than 8.2 kg N/ha/yr for low-elevation
desert containing invasive Mediterranean grass (Schismus barbatus) and less than
5.7 kg N/ha/yr at higher elevation sites containing non-native red brome (Bromus rubens;
Rao et al.. 2010). Invasive non-native grass production in JOTR was simulated under
varying levels of N input and precipitation. Results suggested changing fire frequency.
Simulated fire risk was higher when N deposition was above 3 kg N/ha/yr, but then
stabilized at N deposition above 5.7 kg N/ha/yr in pinyon-juniper and at 8.2 kg N/ha/yr in
creosote bush scrub plant communities (Rao et al.. 2010). Model results of this study also
suggested that fire risk is controlled more by precipitation than by grass productivity
(Allen and Geiser. 2011).
16.6.3.2. Aquatic
This section highlights recent monitoring, dose-response, and critical load studies of N
and S deposition to aquatic systems in SEKI, JOTR, and similar southern California
ecosystems.
16.6.3.2.1. Acidification
In the Sierra Nevada, both S and N deposition can contribute mobile acid anions (SO42 ,
NO3 ) to watershed soil solution, streams, and lakes. Many aquatic and terrestrial
ecosystems in the Sierra Nevada are generally sensitive to acidification. Lakes in these
mountains are highly sensitive to impacts from acidic deposition because of the granitic
bedrock, thin acidic soils, large amounts of precipitation, coniferous trees, and dilute
surface waters (Mclack and Stoddard. 1991; Melacketal.. 1985; McColl. 1981). Because
the levels of acidic deposition at the higher elevations have been relatively low, however,
acidification effects to date have been minimal.
Many lakes in SEKI have received N deposition high enough to cause limited chronic
N03 leaching, which can contribute to both acidification and eutrophication. Widespread
chronic lake or stream acidification has not occurred to any degree, although Sullivan
(2000) concluded that some episodic acidification has likely occurred. Acid anion
June 2018
16-166
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
concentrations (mainly SOr and NOs ) in most high-elevation lakes were low in surveys
conducted during summer and fall, but higher during spring when snowmelt commonly
causes pulses of high NOs concentrations in surface waters (Melack et al.. 1989).
During the fall of 1985, about one-third of the lakes sampled by the Western Lake Survey
(WLS) in and near SEKI had acid neutralizing capacity (ANC) < 50 |icq/L. and two of
the lakes in the park had ANC < 20 (j,eq/L (Landers etal.. 1987). The possibility of
recovery from lake acidification during the period 1985-1999 was evaluated by another
chemical survey conducted by the U.S. Geological Survey (USGS) during the fall of
1999 (Clow and Sueker. 2000). Clow et al. (2003) reported that the resampled lakes were
generally low in ionic strength and had pH between about 6.0 and 7.0; the median ANC
was 59 (j,eq/L in SEKI and 32 |icq/L in YOSE located to the north. The observed lake
chemistry further confirmed the high sensitivity to acidification.
Key acidification characterization and monitoring studies conducted in the Sierra Nevada
are listed in Table 16-32. None of these studies are recent (post-2008) other than the
study of Clow et al. (2010) in YOSE. The current condition of these waters is probably
similar to what was found during previous decades with no additional recovery.
The weight of evidence suggests that many high-elevation lakes in SEKI have in the past
received N deposition high enough to cause some chronic NO3 leaching. This can
contribute to both acidification and eutrophication. Widespread chronic lake or stream
acidification has not occurred to any degree, although Sullivan (2000) concluded that
some episodic acidification has likely occurred. Acid anion concentrations (mainly SO42
and NO3 ) in most high-elevation lakes were low during summer and fall, but higher
during spring when there are commonly pulses of high NO3 concentrations in surface
waters during snowmelt (Melack et al.. 1989). which is seldom sampled due to logistical
and safety constraints. Acid-base chemistry is not expected to have changed appreciably
in recent years.
16-167
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Table 16-32 Example surface water acidification studies in Sequoia and Kings
Canyons National Parks and other Sequoia and Kings Canyons
National Parks—relevant areas in the southern California case study
region.
Study
Location
Time
Period
Focus
Clow et al. (2003) Clow
and Sueker (2000)
National parks in western
U.S., including SEKI
1999
Resurvey of WLS lakes
Clow et al. (2003)
Sierra Nevada region
1985 and
1999
Decrease in ANC in lakes in SEKI partly due
to climatic differences at time of sampling
Clow et al. (2010)
YOSE
2003
Multiple linear regression modeling to
predict NO3" concentrations and ANC in
lakes across the park
ANC = acid neutralizing capacity; N03 = nitrate; SEKI = Sequoia and Kings Canyons National Parks; WLS = Western Lake
Survey; YOSE = Yosemite National Park.
The hydrology of alpine and subalpine ecosystems in the Sierra Nevada is controlled by
snowmelt because more than 90% of the annual precipitation in this region falls as snow.
The relatively small loads of acidic deposition can contribute to moderately high
concentrations of SOr and NOs in lakes and streams during the early phase of
snowmelt (Stoddard. 1995). Potential biological effects of acidic deposition on surface
waters in the Sierra Nevada are most likely caused by acidification attributable to high
N03 concentrations, which tend to be episodic rather than chronic (Wigington ct al..
1990).
Climatic fluctuations that control the amount and timing of precipitation, melting of the
snowpack, growth of plants, depth to groundwater, and concentration by evaporation of
constituents in solution all influence soil and surface water chemistry. These factors
regulate interactions between air pollutants and the aquatic and terrestrial biological
receptors. Because conditions vary over time, many years of data might be required to
establish the existence of trends in episodic surface water chemistry (Sullivan. 2000).
Effects of acidification on vertebrate animals in the Sierra Nevada have not been
documented. However, declines and elimination of frogs and toads due to uncertain
causes have been documented throughout the western U.S., including in the national
parks in California. Possible impacts of air pollution on aquatic amphibians is noteworthy
because amphibians are especially sensitive to environmental change or degradation
(Bradford and Gordon. 1992; Blaustein and Wake. 1990).
June 2018
16-168
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
Arid and semiarid ecosystems in southern California are generally not known to be
sensitive to acidification from either S or N deposition. Precipitation to these ecosystems
is limited. There are relatively few surface waters, and it appears that they have
appreciable acid buffering capacity. However, semiarid chapparall soils under very high
N deposition in the San Gabriel Mountains have decreased considerably in pH between
the 1970s and 1990s (Fenn et al.. 2011a; Pardo et al.. 201 lc).
16.6.3.2.1.1.	Empirical Studies
Empirical studies have shed light on dose-response relationships in lakes and streams of
southern California. Several new studies are addressed here. Heard et al. (2014)
investigated CLs to protect against lake acidification in the Sierra Nevada. An
experimental study by Kratz et al. (1994) investigated responses of macroinvertebrates.
Earlier studies of effects of N deposition in this region (mainly before 2008) addressed
aspects of water chemistry and the response of algae to nutrient addition. More recent
publications include the reviews of Baron et al. (2011a). Pardo etal. (2011c). and Fenn et
al. (2015).
Heard et al. (2014) investigated the likelihood that paleoreconstructed changes in ANC in
lakes of the Sierra Nevada had been caused by atmospheric deposition of SOr and NOs
during the 20th century. The deposition estimates suggested that the CL at Moat Lake
was exceeded in about 1920 and the chemical recovery started in about 1970. The initial
inferred decline in the ANC of Moat Lake occurred between 1920 and 1930. This time
period corresponded with estimated acidic deposition (SO42 + NO3 ) equal to about
74 eq/ha/yr. This was taken (Heard et al.. 2014) to be the critical load (CL) to protect this
lake against acidification. Diatom-reconstructed ANC values changed from near
100 |icq/L before 1920 to near 60 |icq/L during the 1970s. Recovery of ANC after 1970
was attributed by Heard et al. (2014) to decreased deposition of S. In addition, during the
late 20th century, warmer air temperatures may have contributed to higher ANC via
increased weathering rates.
16.6.3.2.1.2.	Modeling Studies
Some limited work has been conducted on modeling of aquatic critical loads using
steady-state approaches for acidity. Data from over 12,500 streams and lakes were used
by the Critical Loads of Atmospheric Deposition (CLAD) science committee of the
NADP (http://nadp. sws .uiuc.edu/committees/clad/) to model steady-state CLs for acidity
of surface waters based on multiple approaches for estimating base cation weathering.
Shaw et al. (2014) used the Steady-State Water Chemistry (SSWC) model to estimate the
June 2018
16-169
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
steady-state CL of acidity for 208 Sierra Nevada lakes. Study lakes were generally dilute
(mean specific conductance 8 (j,S/cm) and had relatively low ANC (mean 57 |icq/L).
Most were located in Forest Service wilderness areas, some in proximity to Sierra
Nevada national parks. Using a critical ANC limit of 10 (j,eq/L to protect aquatic biota
from effects caused by water acidification, the model predicted a CL of 149 eq/ha
(14.9 meq/m2) of acidity for Sierra Nevada watersheds situated on granitic bedrock. More
than one-third of the study lakes received acidic deposition that was in exceedance of the
estimated CL. The median study lake showed a CL exceedance of about 80 eq/ha
(8 meq/m2/year). Based on these CL calculations, Shaw et al. (2014) concluded that
high-elevation lakes in the Sierra Nevada have not fully recovered from effects of
acidifying deposition despite large decreases in S and N air pollution and reduced S
deposition over the last several decades.
Aquatic Nutrient Enrichment
There have been several studies of water chemistry dose-response relationships that
pertain to nutrient enrichment in the Sierra Nevada case study area. These have included
studies of N retention and release, critical loads, and N saturation. A meta-analysis of
lakes from 42 regions of Europe and North America concluded that atmospheric N
deposition has caused higher concentrations of NOs in lake water and increased
phytoplankton biomass (Bcrgstrom and Jansson. 2006). Bergstrom et al. (2005) found N
limitation in lakes that received N deposition below approximately 2.5 kg N/ha/yr,
limitation of N and P at N deposition between -2.5 and 5.0 kg N/ha/yr, and P limitation
in lakes that had N deposition higher than about 5.0 kg N/ha/yr. These findings may be
relevant to remote lakes in the Sierra Nevada, including those in SEKI.
Sickman et al. (2002) reported yields and retention of dissolved inorganic N (DIN) at
28 high-elevation lakes, including those within SEKI. Net DIN retention was
1.2 kg N/ha/yr, an estimated 55% of annual DIN loading. Sickman et al. (2001)
calculated that the annual yield of N from the Emerald Lake watershed varied by a factor
of 8 (0.4 to 3.2 kg N/ha/yr), which explained 89 and 74% of the variation in DIN and
organic N, respectively. The N content of runoff was higher during years that
experienced high runoff and was lower during dry years. Increases in NOs concentration
were larger during years with deep, late-melting snowpacks. Therefore, it is expected that
climate warming, with associated earlier snowmelt, might increase N retention in
high-elevation watersheds (Sickman et al.. 2001).
Nutrient enrichment of Sierra Nevadan lakes is not solely a function of N inputs.
Atmospheric P deposition can also be important. Sickman et al. (2003) proposed that
atmospheric P deposition and enhanced cycling of P due to climate change were the most
16-170
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
likely sources of the observed increase in P loading to the Sierra Nevada lakes. It is not
known why atmospheric deposition of P to these lakes has increased over time.
Possibilities include increased use of organo-phosphate pesticides (Kcglcv et al.. 2000)
and wind-blown transport of soils and dust that are high in P from the agricultural San
Joaquin Valley to the mountains (Sickman et al.. 2003; Lesack and Melack. 1996;
Bergametti etal.. 1992).
Based on results of bioassay experiments and application of coarse nutrient limitation
indices, Sickman et al. (2003) concluded that phytoplankton growth in Emerald Lake in
SEKI during the early 1980s was limited by P supply. During subsequent years, in
response to increased atmospheric P input, the nutrient balance shifted into a pattern of
colimitation by N + P, and finally N limitation. Sickman et al. (2003) observed that the
total P concentration in the lake doubled from 1983 to 1999. Particulate C concentration
in the lake in 1999 was four times higher than the average during the period 1983-1998.
Together with the observed increase in total P in lake water, this suggested that
eutrophication had occurred during that time (Sickman et al.. 2003). It was proposed that
increased P loading to Sierra Nevada lakes decreased NOs concentration in lake water
and promoted N limitation. The increased P loading may have been due to increased
emissions and deposition of P and/or accelerated internal P cycling that might be caused
by changes in climatic conditions and the timing of snowmelt and runoff (Sickman et al..
2003; Johnson. 1998: Dettingerand Cavan. 1995).
Baron etal. (2011a) synthesized CLs of N deposition for protecting against nutrient
enrichment of high-elevation lakes. Relationships between NO3 concentrations and N
deposition suggested a CL near 2 kg N/ha/yr to prevent NO3 leaching to lakes in the
Sierra Nevada. Fennetal. (2011a) also summarized data reflecting the relationship
between atmospheric N deposition and stream NO3 concentration, in this case for
chaparral vegetation in the San Dimas Experimental Forest in the San Gabriel Mountains
(Meixner et al.. 2006; Riggan et al.. 1994; Riggan et al.. 1985). Devil Canyon watershed
in the western San Bernardino Mountains (Meixner and Fenn. 2004; Fenn and Poth.
1999). and Chamise Creek in SEKI (Fenn et al.. 2003a; Fenn et al.. 2003c). Stream water
N03 concentrations in streams draining chaparral vegetation were among the highest
reported in North America, with concentrations in some streams higher than 200 |icq/L
(Fennetal.. 2011a; Fenn et al.. 2003a; Fenn et al.. 2003c; Fenn and Poth. 1999; Riggan et
al.. 1994; Riggan et al.. 1985). Evaluation of N saturation in chaparral must also consider
the prevalence of fire in this ecosystem type because the stand-replacing fire interval is
only about 40 to 60 years (Minnichand Bahre. 1995). Much of the atmospherically
deposited N that has accumulated in the litter and aboveground vegetation over the
intervening years is released suddenly to the atmosphere during fire. In contrast, much of
the accumulated N in the mineral soil generally remains on-site after a fire (Fenn et al..
16-171
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
201 la), and fire contributes to increased NO3 concentrations in stream water (Riggan et
al.. 1994; Anderson and Poth. 1989).
The recent work that has been conducted on aquatic ecosystems has focused mostly on
the Sierra Nevada region. Analogous nutrient enrichment work has not been conducted in
or around JOTR, in part, because there are few surface waters in that region.
16.6.3.2.2.1. Algae
Changes in nutrient supply can affect aquatic biota at all trophic levels, but algae are
perhaps most likely to show eutrophication effects from N deposition. Studies have
shown an increase in lake phytoplankton biomass in response to increases in N deposition
in the Sierra Nevada region (Sickman et al.. 2003). Wyoming (Lafrancois et al.. 2003b').
Sweden (Bcrgstrom et al.. 2005). and across Europe (Bcrgstrom and Jansson. 2006).
Increases in algal biomass have been associated with changes in algal assemblages that
favor certain species over others. A widespread increase in the relative abundance of
Asterionella formosa and Fragilaria crotonensis occurred in oligotrophic lakes across the
western U.S.; these changes have been documented from both lake sediment cores and
limnological surveys (Saros et al.. 2003; Wolfe et al.. 2001; Interlandi et al.. 1999;
Goldman. 1988). In Lake Tahoe to the north of YOSE, there has been an increase since
about 1950 in the ratio of araphidinate pennate diatoms to centric diatoms. This was
largely due to increased abundance of Fragilaria crotonensis. This change in diatom
relative abundance was associated with higher N loading to the lake (Goldman. 1988).
Jassbv et al. (1994) showed that atmospheric deposition supplies most of the N to Lake
Tahoe.
Because high-elevation lakes in the Sierra Nevada tend to be highly oligotrophic, small
changes in nutrient supply can affect algal productivity (Sickman et al.. 2003). Chamise
Creek in SEKI was reported to have high NO3 leaching in response to throughfall N
deposition near 10 kg N/ha/yr, suggesting N saturation (Fenn et al.. 2003a; Fenn et al..
2003c). The U.S. Forest Service has suggested a policy threshold of 2 |icq/L for stream
N03 concentration in N limited ecosystems in the western U.S. This load has been
suggested as a tipping point to trigger management concern for possible over-enrichment
of aquatic ecosystems with N (Fenn et al.. 2011b).
16.6.4. Highlights of Additional Research Literature and Federal Reports since
January 2008
Key research literature published since January 2008 is highlighted in Table 16-33.
June 2018	16-172	DRAFT: Do Not Cite or Quote

-------
Table 16-33 Key recent research literature focused on the case study region.
Publication
Focus
Allen and Geiser(2011)
N addition experiment in JOTR
Allen et al. (2009)
Soil N along depositional gradient in JOTR
Baron et al. (2011a)
Critical load and dose-response review
Bowman et al. (2011)
Critical loads for high-elevation vegetation
Bvtnerowicz et al. (2015)
Empirical N critical loads in San Bernardino Mts.
Clow et al. (2010)
Lake acid-base chemistry characterization in YOSE
Clark etal. (2013)
Loss of herbaceous plant species in response to N deposition
Cox etal. (2014)
Conversion from coastal sage scrub to grassland
Ellis etal. (2013)
Empirical critical loads
Esque et al. (2010)
N dynamics after fire in Mojave Desert shrubs
Fenn et al. (2008)
Effects of N on epiphytic lichens and stream nitrate leaching
Fenn etal. (2010)
Empirical critical loads and exceedances
Fenn etal. (2011a)
Critical loads review
Fenn etal. (2015)
Critical loads review
Geiser et al. (2010).
Effects of N on lichens
Grulke et al. (2008)
Forest susceptibility to wildfire in San Bernardino Mts.
Heard etal. (2014)
Critical loads for lakes
Hurteau and North (2009)
Effects of N on a sensitive herbaceous plant species
Johnson et al. (2011)
Soil acid-base chemistry in mixed conifer watersheds
Jovan (2008)
Response of lichens to N
Kimball etal. (2014)
Effects of water and N on CSS following fire
McCallev and Sparks (2009)
Soil temperature effects on N loss in Mojave Desert
Pardo et al. (2011c)
Critical load and exceedance for nutrient N enrichment
Pasauini and Vourlitis (2010)
Net primary production in chaparral
Rao et al. (2009)
Effects of N deposition on mineralization
June 2018
16-173 DRAFT: Do Not Cite or Quote

-------
Table 16-33 (Continued): Key recent research literature focused on the case study
region.
Publication
Focus
Rao et al. (2010)
N and fire effects on semiarid plant communities
Rao et al. (2011)
Critical loads for desert vegetation
Shaw et al. (2014)
Critical loads for lakes in Sierra Nevada
Stark etal. (2011)
Effects of N deposition and climate on Mojave Desert vegetation
Talluto and Sudina (2008).
Conversion of CSS to grassland
Vamstad and Rotenberrv (2010)
Changes in plant species composition after fire
Vourlitis (2012)
Response of semiarid vegetation to N and climate
Vourlitis and Fernandez (2012)
N response to N input in semiarid shrublands
Vourlitis and Pasquini (2008)
C and N dynamics in pre- and post-fire chaparral
Vourlitis and Pasauini (2009)
Dry season addition of N to CSS and chaparral
C = carbon; CSS = coastal sage scrub; JOTR = Joshua Tree National Park; Mts = mountains; N = nitrogen; YOSE = Yosemite
National Park.
16.6.5. Summary
1	This case study focuses on the ecosystems found in two national parks in southern
2	California, SEKI and JOTR. SEKI largely consists of forested and mountainous terrain in
3	and near the Sierra Nevada Mountains. JOTR is covered largely by desert and semiarid
4	land. Both are generally downwind of air pollution (mainly N) sources. The majority of
5	the information provided in this case study is applicable to ecosystems found within
6	SEKI. Limited research conducted in JOTR was found in the peer-reviewed literature. A
7	summary of studies relevant to evaluating CLs is shown by Figure 16-54.
June 2018
16-174
DRAFT: Do Not Cite or Quote

-------
m
on
c
"o
Q.
~a
c
<
>-
"a
a
| Manne West Coast Forests-Lichen protection (wet deposttio«)-Geiser et a!., 2010
JOTR-Wetted areas with tow day content-Rao et al.f 2010
| Sierra Nevada Mtns.-Prevent N03- leaching-Fenn et al.t 2011a
JOTR-Creosote bush scrub-Fenn et at., 2008,201S
SEKI and YOSE-Lichen protection-Ellrc et a!.» 2013; Pardo et al., 2011a
California-Protection of 7 vegetation types-Fenn et al,, 2010
Sierra Nevada, mainly SŁKI-Protecnon of mycorrb«al fungi, lichens, herbaceous plants and forest vegetation
and to limit N03- leaching-Pando et al, 2011b
I JOTR-Lichen and herbaceous plants-Pardo et al. 2011b
| Marine Wesi Coast Forests-Lichen protection (total deposition!-Geiser
et al., 2010
Western Sierra Nevada-Epiphytic lichen shift-Bowman et at, 2011. Fenn et al., 2008
San Bemadino Mtns.-Mtxed conifer forest-8ynerowicz et al., 201S
JOTR-Pinyon-juniper-Fenn et al., 2008, 2015
| San Bemadino Mtns.-Lie hen protection-Bynerowicz et al.. 201S
San Bemadino Mtns.-Lichen protection in chapparal ecosystems-Fenn et al., 2010
Calrfornia-Lichen protection; Lichen presence at 3 of S3 lichen survey sites et al., 2011a
JOTR-Higher elevation sites containing non-native red brome-Rao et al., 2010
JOTR-Pinyon-juniper protection-Rao et al., 2010
I SEKMncreased the losses of N as NOB- in streams and gaseous NO in chapparal or
similar semi-and vegetation-ti et al., 2006
~
~
~
~
~
JOTR-lnvasive grasses-AJIen et al., 2009; Allen and Geiser, 2011
JOTR-Creosote bush scrub communities-Rao et al., 2010
JOTR-Low elevation desert containing invasive Mediterranean grass
Raoetal., 2010
~ N. American deserts-Fire risk in I vegetation type-Pardo et al.,
2011a, 2011b
~
N. Amencan deserts-Fire nsk in 1 vegetation type-Pardo et
al. 2011a, 2011b
~ Western Sierra Nevada Oligotrophy lichen elimination
Bowman et al., 2011
¦ So. California-Coastal sage scrub-Cox
et al., 2014
~ So. California -Conversion from coastal sage
scrub to annual grassland-Cox et al, 2014
YOSE-Herbaceous species-Hurteau and North, 2009
~
Semi-arid shrubland-Shrub growth and relative abundance of ["
sub-dominant shrubs and herbaceous plants-Pasquini and L_
Vourlitts, 2010
San Bemadino Mtns.-N03- leaching in chapparal
Fenn et al., 2010
~
San Bemadino MIns.-Protection against N03- leaching in
mixed conifer forests-Bynerowia et al.f 201S
Sierra Nevada and San Bemadino Mtns.-Protection against
N03- leaching Fenn et al.. 2008,201S
So. California-Mixed conifer forest-Breiner, 2007
JOTR-Creosote bush-Rao et al., 2010
JOTR-Pin yon-juniper-Rao et al.,
2010
~5	7	g	5	nr
N deposition, Kg N/ha/yr
~TI	IT
TT
14 >14
ha = hectare; JOTR = Joshua Tree National Park; kg = kilogram; Mtns. = mountains; N = nitrogen; N. = north; NO = nitric oxide;
N03" = nitrate; SEKI = Sequoia and Kings Canyons National Parks; So = southern; YOSE = Yosemite National Park; yr = year.
Figure 16-54
Continuum of critical loads in southern California case study area
and relevant surrounding region.
June 2018
16-175
DRAFT: Do Not Cite or Quote

-------
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
The Sierra Nevada Mountains, in which SEKI is located, contain many acid-sensitive
aquatic resources. Limited monitoring studies indicated that the current condition of these
waters is probably similar to that found during previous decades with no additional
recovery. Aquatic acidification effects to date on SEKI's condition have been minimal at
higher elevations due to low acidic deposition, and widespread lake or stream
acidification has not occurred to any degree, although some episodic acidification has
likely occurred. There are expectations for N retention in high-elevation watersheds as
snowmelt changes temporally due to climate warming. Small changes in nutrient supply
to high-elevation, oligotrophic lakes in the Sierra Nevada can affect algal productivity
and increase abundance of Asterionellaformosa and Fragilaria crotonensis (Saros ct al..
2003). Amphibian decline and the possible association with air pollution is noteworthy.
However, one of the most pronounced effects of N deposition in the Sierra Nevada has
been shown to be an alteration of the lichen community which has CLs ranging from 3.1
to 5.2 kg N/ha/yr (Bowman et al.. 2011; Fenn et al.. 2008).
JOTR's arid and semiarid ecosystems are not known to be sensitive to acidification
because there is limited precipitation, few surface waters, and high acid buffering
capacity. N deposition may increase productivity of invasive grasses after fire, and
deposition and climate change may promote an altered fire cycle that does not allow
sufficient time for Joshua tree woodlands to re-establish between fires. It has been
suggested that N deposition may have already exceeded JOTR's critical load of
8.4 kg N/ha/yr for terrestrial plant communities (Pardo et al.. 2011c).
16-176
DRAFT: Do Not Cite or Quote

-------
References
Abdel-Halim. KY; Abo El-Saad. AM; Talha. MM; Hussein. AA; Bakry. NM. (2013). Oxidative
stress on land snail Helix aspersa as a sentinel organism for ecotoxicological effects of
urban pollution with heavy metals. Chemosphere 93: 1131-1138.
http://dx.doi. org/10.1016/i. chemosphere.2013.06.042.
Abel 1, GCJ; Revill, AT; Smith. C; Bissett AP; Volkman. JK; Robert. SS. (2010). Archaeal
ammonia oxidizers and nirS-type denitrifiers dominate sediment nitrifying and
denitrifying populations in a subtropical macrotidal estuary. ISME J 4: 286-300.
http://dx.doi.org/10.1038/ismei.20Q9.105.
Abell. RA; Loucks. CJ; Hedao. P; Olson. DM; Dinerstein. E; Hurley. P; Diggs. JT; Eichbaum.
W; Walters, S; Wettengel, W; Allnutt, T. (1999). Freshwater ecoregions of North
America : A conservation assessment. In Freshwater Ecoregions of North America : A
Conservation Assessment. Washington, D.C.: Island Press.
http://www.islandpress.Org/ip/books/book/islandpress/F/bo3559448.html.
Aber, J; McDowell, W; Nadelhoffer, K; Magill, A; Berntson, G; Kamakea, M; McNulty, S;
Currie. W; Rustad. L; Fernandez. I. (1998). Nitrogen saturation in temperate forest
ecosystems. Bioscience 48: 921-934. http://dx.doi.org/10.2307/1313296.
Aber, JD; Federer, CA. (1992). A generalized, lumped-parameter model of photosynthesis,
evapotranspiration and net primary production in temperate and boreal forest ecosystems.
Oecologia 92: 463-474. http://dx.doi.org/10.1007/BF0Q317837.
Aber, JD; Goodale, CL; Ollinger, SV; Smith, ML; Magill, AH; Martin, ME; Hall, RA; Stoddard,
JL. (2003). Is nitrogen deposition altering the nitrogen status of northeastern forests?
Bioscience 53: 375-389.
Aber, JD; Magill, A; Mcnultv, SG; Boone, RD; Nadelhoffer, KJ; Downs, M; Hallett, R. (1995).
Forest biogeochemistry and primary production altered by nitrogen saturation. Water Air
Soil Pollut 85: 1665-1670. http://dx.doi.org/10.1007/BF0Q477219.
Aber, JD; Nadelhoffer, KJ; Steudler, P; Melillo, JM. (1989). Nitrogen saturation in northern
forest ecosystems: Excess nitrogen from fossil fuel combustion may stress the biosphere.
Bioscience 39: 378-386. http://dx.doi.org/10.2307/1311067.
Aber, JD; Ollinger, SV; Driscoll, C. (1997). Modeling nitrogen saturation in forest ecosystems in
response to land use and atmospheric deposition. Ecol Modell 101: 61-78.
http://dx.doi.org/10.1016/S0304-3800(97)01953-4.
Aber, JD; Ollinger, SV; Driscoll, CT; Likens, GE; Holmes, RT; Freuder, RJ; Goodale, CL.
(2002). Inorganic nitrogen losses from a forested ecosystem in response to physical,
chemical, biotic, and climatic perturbations. Ecosystems 5: 648-658.
http://dx.doi. org/10.1007/s 10021 -002-0203 -8.
Abreu, MH; Pereira, R; Buschmann, AH; Sousa-Pinto, I; Yarish, C. (2011). Nitrogen uptake
responses of Gracilaria vermiculophylla (Ohmi) Papenfuss under combined and single
addition of nitrate and ammonium. Exp Mar Bio Ecol 407: 190-199.
http://dx.doi.Org/10.1016/i.iembe.2011.06.034.
June 2018
R-l
DRAFT: Do Not Cite or Quote

-------
ABS. P; Muniz, IP. (1993). The effect of aluminium in soft water at low pH and different
temperatures on mortality, ventilation frequency and water balance in smoltifying
Atlantic salmon, Salmo salar. Environmental Biology of Fishes 36: 193-203.
http://dx.doi.org/10.1007/BF000Q2799.
Acha. D; Hintelmann. H: Yee. J. (2011). Importance of sulfate reducing bacteria in mercury
methylation and demethylation in periphyton from Bolivian Amazon region.
Chemosphere 82: 911-916. http://dx.doi.Org/10.1016/i.chemosphere.2010.10.050.
Ackerman. JT; Eagles-Smith. CA. (2010). Agricultural wetlands as potential hotspots for
mercury bioaccumulation: Experimental evidence using caged fish. Environ Sci Technol
44: 1451-1457. http://dx.doi.org/10.1021/es9Q28364.
Ackerman. JT: Miles. AK; Eagles-Smith. CA. (2010). Invertebrate mercury bioaccumulation in
permanent, seasonal, and flooded rice wetlands within California's Central Valley. Sci
Total Environ 408: 666-671. http://dx.doi.Org/10.1016/i.scitotenv.2009.10.030.
Adams. KE; Taranu. ZE; Zurawell. R; Cumming. BF; Gregory-Eaves. I. (2014). Insights for lake
management gained when paleolimnological and water column monitoring studies are
combined: A case study from Baptiste Lake. Lake Reserv Manag 30: 11-22.
http://dx.doi.org/10.1080/10402381.2Q13.865687.
Adams. MB: Angradi. TR; Kochenderfer. JN. (1997). Steam water and soil solution responses to
5 years of nitrogen and sulfur additions at the Fernow Experimental Forest, West
Virginia. For Ecol Manage 95: 79-91. http://dx.doi.org/10.1016/S0378-l 127(96^)03990-4.
Adams, MB; Burger, JA; Jenkins, AB; Zelazny, L. (2000). Impact of harvesting and atmospheric
pollution on nutrient depletion of eastern US hardwood forests. For Ecol Manage 138:
301-319. http://dx.doi.org/10.1016/S0378-l 127(0QN)00421-7.
Adams, MB: Nichols, D; Feder, CA: Jensen, K; H, P. (1991). Screening procedure to evaluate
effects of air pollution on eastern region wildernesses cited as Class I Air Quality areas.
(NE-151). Radnor, PA: U.S. Department of Agriculture, Forest Service, Northeastern
Forest Experiment Station.
Adrian, R; O'Reilly, CM: Zagarese, H; Baines, SB: Hessen, D, agO; Keller, W: Livingstone,
DM; Sommaruga, R; Straile, D; Van Donk, E; Weyhenmeyer, GA; Winder, M. (2009).
Lakes as sentinels of climate change. Limnol Oceanogr 54: 2283-2297.
Aerts, R. (1996). Nutrient resorption from senescing leaves of perennials: Are there general
patterns? J Ecol 84: 597-608. http://dx.doi.org/10.2307/2261481.
Aerts, R. (1997). Climate, leaf litter chemistry and leaf litter decomposition in terrestrial
ecosystems: A triangular relationship. Oikos 79: 439-449.
http://dx.doi.org/10.2307/3546886.
Aerts, R. (2009). Nitrogen supply effects on leaf dynamics and nutrient input into the soil of
plant species in a sub-arctic tundra ecosystem. Polar Biology 32: 207-214.
http://dx.doi.org/10.1007/s00300-008-Q521-l.
Aerts, R; Berendse, F; de Caluwe, H; Schmitz, M. (1990). Competition in heathland along an
experimental gradient of nutrient availability. Oikos 57: 310-318.
http://dx.doi.org/10.2307/3565959.
June 2018
R-2
DRAFT: Do Not Cite or Quote

-------
Aerts, R; Chapin, FS, III. (1999). The mineral nutrition of wild plants revisited: A re-evaluation
of processes and patterns. Adv Ecol Res 30: 1-67. http://dx.doi.org/10.1016/S0Q65-
2504(08)60016-1.
Aerts, R; de Caluwe, H. (1999). Nitrogen deposition effects on carbon dioxide and methane
emissions from temperate peatland soils. Oikos 84: 44-54.
http://dx.doi.org/10.2307/3546865.
Aguillaume. L; Izquieta-Roiano. S: Garcia-Gomez. H; Elustondo. D; Santamaria. JM: Alonso. R;
Avila. A. (2017). Dry deposition and canopy uptake in Mediterranean holm-oak forests
estimated with a canopy budget model: A focus on N estimations. Atmos Environ 152:
191-200. http://dx.doi.Org/10.1016/i.atmosenv.2016.12.038.
Aherne. J: Farrell. EP. (2002). Deposition of sulphur, nitrogen and acidity in precipitation over
Ireland: Chemistry, spatial distribution and long-term trends. Atmos Environ 36: 1379-
1389. http://dx.doi.org/10.1016/S1352-2310(01)00507-6.
Aherne. J: Larssen, T; Dillon. PJ; Cosby. BJ. (2004). Effects of climate events on elemental
fluxes from forested catchments in Ontario, Canada: Modelling drought-induced redox
processes. Water Air Soil Pollut 4: 37-48.
http://dx.doi.Org/10.1023/B:WAFO.0000028343.23645.93.
Aherne. J: Posch. M. (2013). Impacts of nitrogen and sulphur deposition on forest ecosystem
services in Canada. Current Opinion in Environmental Sustainability 5: 108-115.
http://dx.doi.Org/10.1016/i.cosust.2013.02.005.
Ainsworth, CH; Samhouri, JF; Busch, DS; Cheung, WWL; Dunne, J; Okey, TA. (2011).
Potential impacts of climate change on Northeast Pacific marine foodwebs and fisheries.
ICES J Mar Sci 68: 1217-1229. http://dx.doi.org/10.1093/icesims/fsr043.
Akerblom, S: Baath, E; Bringmark, L; Bringmark, E. (2007). Experimentally induced effects of
heavy metal on microbial activity and community structure of forest mor layers. Biol
Fertil Soils 44: 79-91. http://dx.doi.org/10.1007/sQ0374-007-0181-2.
Akerblom, S: Bishop, K; Biorn, E; Lambertsson, L; Eriksson, T; Nilsson, MB. (2013).
Significant interaction effects from sulfate deposition and climate on sulfur
concentrations constitute major controls on methylmercury production in peatlands.
Geochim Cosmo Act 102: 1-11. http://dx.doi.Org/10.1016/i.gca.2012.10.025.
Al-Reasi, HA: Wood, CM: Smith, DS. (2013). Characterization of freshwater natural dissolved
organic matter (DOM): Mechanistic explanations for protective effects against metal
toxicity and direct effects on organisms. Environ Int 59: 201-207.
http://dx.doi.Org/10.1016/i.envint.2013.06.005.
A1 Jawaheri, R; Sahlen, G. (2017). Negative impact of lake liming programmes on the species
richness of dragonflies (Odonata): a study from southern Sweden. Hydrobiologia 788:
99-113. http://dx.doi.Org/l0.1007/s 10750-016-2990-5.
June 2018
R-3
DRAFT: Do Not Cite or Quote

-------
Alcamo. JD; Van Vuuren. DP; Cramer. W; Alder. J; Bennett EM; Carpenter. S; Christen sen. V;
Foley. J; Maerker. M; Masui. T; Morita. T; O'Neill. B; Peterson. G; Ringler, C;
Rosegrant, M; Schulze, K; Bouwman, L; Eickhout, B; Martina, F; Lai, R; Takahashi, K.
(2005). Changes in ecosystem services and their drivers across the scenarios. In A
Millennium Ecosystem (Ed.), Millennium Ecosystem Assessment: Ecosystems and
Human Weil-Being: Scenarios (pp. 297-373). Washington, D.C.: Island Press.
https://www.researchgate.net/publication/235866401 Changes in ecosystem services a
nd their drivers across the scenarios.
Aldous, AR. (2002a). Nitrogen retention by Sphagnum mosses: Responses to atmospheric
nitrogen deposition and drought. Can J Bot 80: 721-731. http://dx.doi.org/10.1139/BQ2-
054.
Aldous, AR. (2002b). Nitrogen translocation in Sphagnum mosses: Effects of atmospheric
nitrogen deposition. NewPhytol 156: 241-253. http://dx.doi.org/10.1046/i. 1469-
8137.2002.00518.x.
Ale, MT; Mikkelsen, JD; Meyer, AS. (2011). Differential growth response of Ulva lactuca to
ammonium and nitrate assimilation. J Appl Phycol 23: 345-351.
http://dx.doi.org/10.1007/slQ811-010-9546-2.
Alexander, RB; Bover, EW; Smith, RA; Schwarz, GE; Moore, RB. (2007). The role of
headwater streams in downstream water quality. J Am Water Resour Assoc 43: 41-59.
http://dx.doi.org/10.111 l/i.l752-1688.2007.00005.x.
Alexander, RB; Johnes, RJ; Bover, EW; Smith, RA. (2002). A comparison of models for
estimating the riverine export of nitrogen from large watersheds. Biogeochemistry 57:
295-339. http://dx.doi.Org/10.1023/A:1015752801818.
Alexander, RB; Smith, RA; Schwarz, GE. (2000). Effect of stream channel size on the delivery
of nitrogen to the Gulf of Mexico. Nature 403: 758-761.
http://dx.doi.org/10.1038/35001562.
Alexander, RB; Smith, RA; Schwarz, GE; Bover, EW; Nolan, JV; Brakebill, JW. (2008).
Differences in phosphorus and nitrogen delivery to the Gulf of Mexico from the
Mississippi River Basin. Environ Sci Technol 42: 822-830.
http://dx.doi.org/10.1021/es0716103.
Alexandre, A, na; Silva, J; Buapet, P; Biork, M; Santos, R, ui. (2012). Effects of C02
enrichment on photosynthesis, growth, and nitrogen metabolism of the seagrass Zostera
noltii. 2: 2620-2630. http://dx.doi.org/10.1002/ece3.333.
Alkhatib, M; del Giorgio, PA; Gelinas, Y; Lehmann, MF. (2013). Benthic fluxes of dissolved
organic nitrogen in the lower St. Lawrence estuary and implications for selective organic
matter degradation. Biogeosciences 10: 7609-7622. http://dx.doi.org/10.5194/bg-lQ-
7609-2013.
Allen, EB; Egerton-Warburton, LM; Hilbig, BE; Valliere, JM. (2016). Interactions of arbuscular
mycorrhizal fungi, critical loads of nitrogen deposition, and shifts from native to invasive
species in a southern California shrubland. Botany 94: 425-433.
http://dx.doi.Org/10.l 139/cib-2015-0266.
June 2018
R-4
DRAFT: Do Not Cite or Quote

-------
Allen. EB; Geiser. LH. (2011). North American deserts. In Assessment of Nitrogen Deposition
Effects and Empirical Critical Loads of Nitrogen for Ecoregions of the United States (pp.
133-142). (NRS-80). Newtown Square, PA: U.S. Department of Agriculture, Forest
Service, Northern Research Station. http://www.treesearch.fs.fed.us/pubs/3 8123.
Allen. EB; Padgett PE; Bytnerowicz. A; Minich. R. (1998). Nitrogen deposition effects on
coastal sage vegetation of southern California. In Proceedings of the International
Symposium on Air Pollution and Climate Change Effects on Forest Ecosystems. Albany,
CA: U.S. Department of Agriculture, Forest Service, Pacific Southwest Research Station.
http://www.treesearch.fs.fed.us/pubs/26950.
Allen. EB: Rao. LE; Steers. RJ; Bytnerowicz. A: Fenn. ME. (2009). Impacts of atmospheric
nitrogen deposition on vegetation and soils at Joshua Tree National Park. In RH Webb
(Ed.), The Mojave Desert: Ecosystem Processes and Sustainability (pp. 78-100). Las
Vegas, NV: University of Nevada Press, http://www.treesearch.fs.fed.us/pubs/37082.
Allen. MF; Allen. EB: Lansing. JL; Pregitzer. KS: Hendrick. R. onL; Ruess. RW: Collins. SL.
(2010). Responses to chronic N fertilization of ectomycorrhizal pinon but not arbuscular
mycorrhizal juniper in a pinon-juniper woodland. J Arid Environ 74: 1170-1176.
http://dx.doi.Org/10.1016/i.iaridenv.2010.05.001.
Allison. SD; Czimczik. CI: Treseder. KK. (2008). Microbial activity and soil respiration under
nitrogen addition in Alaskan boreal forest. Global Change Biol 14: 1156-1168.
http://dx.doi.org/10.1111/i. 1365-2486.2008.01549.x.
Allison. SD: Gartner. TB; Mack. MC: Mcguire. K; Treseder. K. (2010). Nitrogen alters carbon
dynamics during early succession in boreal forest. Soil Biol Biochem 42: 1157-1164.
http://dx.doi.Org/10.1016/i.soilbio.2010.03.026.
Almand-Hunter. BB; Walker. JT; Masson. NP; Hafford. L; Hannigan. MP. (2015). Development
and validation of inexpensive, automated, dynamic flux chambers. Atmos Meas Tech 8:
267-280. http://dx.doi.org/10.5194/amt-8-267-2015.
Almas. AR; Bakken. LR; Mulder. J. (2004). Changes in tolerance of soil microbial communities
in Zn and Cd contaminated soils. Soil Biol Biochem 36: 805-813.
http://dx.doi.Org/10.1016/i.soilbio.2004.01.010.
Almendras. ML: Carballa. M; Diels. L; Vanbroekhoven. K; Chamy. R. (2009). Prediction of
heavy metals mobility and bioavailability in contaminated soil using sequential extraction
and biosensors. J Environ Eng 135: 839-844.
Alonso, A; Camargo, JA. (2013). Nitrate causes deleterious effects on the behaviour and
reproduction of the aquatic snail Potamopyrgus antipodarum (Hydrobiidae, Mollusca).
Environ Sci Pollut Res Int 20: 5388-5396. http://dx.doi.org/10.1007/sl 1356-013-1544-x.
Alpers. CN: Fleck. JA: Marvin-Dipasquale. M; Strieker. CA: Stephenson. M; Taylor. HE.
(2014). Mercury cycling in agricultural and managed wetlands, Yolo Bypass, California:
Spatial and seasonal variations in water quality. Sci Total Environ 484: 276-287.
http://dx.doi.Org/10.1016/i.scitotenv.2013.10.096.
Altieri. AH; Gedan. KB. (2015). Climate change and dead zones [Review], Global Change Biol
21: 1395-1406. http://dx.doi.Org/10.l 111/gcb. 12754.
June 2018
R-5
DRAFT: Do Not Cite or Quote

-------
Altieri, KE; Turpin. BJ; Seitzinger, SP. (2009a). Composition of dissolved organic nitrogen in
continental precipitation investigated by ultra-high resolution FT-ICR mass spectrometry.
Environ Sci Technol 43: 6950-6955. http://dx.doi.org/10.1021/es90Q7849.
Altieri. KE: Turpi n. BJ: Seitzinger. SP. (2009b). Oligomers, organosulfates, and nitrooxy
organosulfates in rainwater identified by ultra-high resolution electrospray ionization FT-
ICR mass spectrometry. Atmos Chem Phys 9: 2533-2542. http://dx.doi.org/10.5194/acp-
9-2533-2009.
Altman. JC: Paerl. HW. (2012). Composition of inorganic and organic nutrient sources
influences phytoplankton community structure in the New River Estuary, North Carolina.
Aquatic Ecology 46: 269-282. http://dx.doi.org/10.1007/slQ452-012-9398-8.
Alvarado. MJ; Cadv-Pereira. KE: Xiao. Y; Millet DB; Payne. VH. (2011). Emission ratios for
ammonia and formic acid and observations of peroxy acetyl nitrate (PAN) and ethylene
in biomass burning smoke as seen by the tropospheric emission spectrometer (TES).
Atmosphere (Basel) 2: 633-654. http://dx.doi.org/10.3390/atmos2040633.
Alvo. R. (2009). Common loon, Gavia immer, breeding success in relation to lake pH and lake
size over 25 years. Canadian Field-Naturalist 123: 146-156.
AM EC Environment & Infrastructure, Inc., . (2015). Clean air status and trends network
(CASTNET) 2013 annual report, EPA contract no. EP-W-15-003. Washington, DC: U.S.
Environmental Protection Agency, Office of Air and Radiation, Clean Air Markets
Division, http://epa.gov/castnet/iavaweb/docs/annual report 2013.pdf.
Ammann, M; Von Ballmoos, P; Stalder, M; Suter, M; Brunold, C. (1995). Uptake and
assimilation of atmospheric N02-N by spruce needles (Picea abies): A field study. Water
Air Soil Pollut 85: 1497-1502. http://dx.doi.org/10.1007/BF0Q477193.
Amthor, JS. (2000). The McCree-de Wit-Penning de Vries-Thornley respiration paradigms: 30
years later. Ann Bot 86: 1-20. http://dx.doi.Org/10.1006/anbo.2000.l 175.
Amundson, RG: MacLean, DC. (1982). Influence of oxides of nitrogen on crop growth and
yield: An overview. In T Schneider; L Grant (Eds.), Studies in Environmental Science
21: Air pollution by nitrogen oxides: Proceedings of the US-Dutch international
symposium, Maastricht, The Netherlands, May 24-28, 1982 (pp. 501-510). Amsterdam,
The Netherlands: Elsevier. http://dx.doi.org/10.1016/B978-0-444-42127-2.5005Q-6.
Anas, MUM: Scott, KA; Cooper, RN; Wissel, B. (2014). Zooplankton communities are good
indicators of potential impacts of Athabasca oil sands operations on downwind boreal
lakes. Can J Fish Aquat Sci 71: 719-732. http://dx.doi.org/10.1139/cifas-2013-0472.
Andersen, DK; Ejrnaes, R; Riis, T. (2016). N- and P-addition inhibits growth of rich fen
bryophytes. Journal of Bryology 38: 127-137.
http://dx.doi.org/10.1080/03736687.2Q15.1114695.
Andersen, JH; Fossing, H; Hansen, JW: Manscher, OH: Murray, C: Petersen, PL. (2014).
Nitrogen inputs from agriculture: Towards better assessments of eutrophication status in
marine waters. Ambio 43: 906-913. http://dx.doi.org/10.1007/sl3280-014-Q514-v.
June 2018
R-6
DRAFT: Do Not Cite or Quote

-------
Andersen. JH; Schluter. L; Aertebierg. G. (2006). Coastal eutrophication: recent developments in
definitions and implications for monitoring strategies. Journal of Plankton Research
28: 621-628.
Andersen. KM: Corre. MP: Turner. BL; Palling. JW. (2010). Plant-soil associations in a lower
montane tropical forest: Physiological acclimation and herbivore-mediated responses to
nitrogen addition. Funct Ecol 24: 1171-1180. http://dx.doi.Org/10.l 11 l/j.1365-
2435.2010.01731.x.
Anderson. PC: Loughner. CP: Piskin. G: Weinheimer. A: Canty. TP: Salawitch. RJ; Worden.
HM; Fried. A: Mikoviny. T; Wisthaler. A: Pickerson. RR. (2014a). Measured and
modeled CO and NOy in PISCOVER-AQ: An evaluation of emissions and chemistry
over the eastern US. Atmos Environ 96: 78-87.
http://dx.doi.Org/10.1016/i.atmosenv.2014.07.004.
Anderson. IC: Brush. MJ: Piehler. MF: Currin. CA: Stanhope. JW: Smyth. AR; Maxev. JP:
Whitehead. ML. (2014b). Impacts of Climate-Related Privers on the Benthic Nutrient
Filter in a Shallow Photic Estuary. Estuaries Coasts 37: S46-S62.
http://dx.doi.org/10.1007/sl2237-013-9665-5.
Anderson. IC: Poth. MA. (1989). Semiannual losses of nitrogen as NO and N20 from unburned
and burned chaparral. Global Biogeochem Cycles 3: 121-135.
http://dx.doi.org/10.1029/gb003i002pQ0121.
Anderson. JAH: Hooper. MJ: Zak. JC: Cox. SB. (2009a). Characterization of the structural and
functional diversity of indigenous soil microbial communities in smelter-impacted and
nonimpacted soils. Environ Toxicol Chem 28: 534-541. http://dx.doi.org/10.1897/Q8-
281.1.
Anderson. JAH: Hooper. MJ: Zak. JC: Cox. SB. (2009b). Molecular and functional assessment
of bacterial community convergence in metal-amended soils. Microb Ecol 58: 10-22.
http://dx.doi.org/10.1007/s00248-0Q8-9467-7.
Anderson. TR: Groffman. PM: Walter. MT. (2015). Using a soil topographic index to distribute
denitrification fluxes across a northeastern headwater catchment. J Hydrol 522: 123-134.
http://dx.doi.Org/10.1016/i.ihydrol.2014.12.043.
Andersson. AJ: Kline. PI; Edmunds. PJ: Archer. SP: Bednarsek. N: Carpenter. RC: Chadsev. M.
eg; Goldstein. P; Grottoli. AG; Hurst. TP; King. AL; Kuebler. JE; Kuffner. IB; Mackev.
KRM; Menge. BA; Pavtan. A; Riebesell. U. If; Schnetzer. A; Warner. ME; Zimmerman.
RC. (2015). Understanding ocean acidification impacts on organismal to ecological
scales. Oceanography 28: 16-27.
Andren. CM; Rydin. E. (2012). Toxicity of inorganic aluminium at spring snowmelt—In-stream
bioassays with brown trout (Salmo trutta L.). Sci Total Environ 437: 422-432.
http://dx.doi.Org/10.1016/i.scitotenv.2012.08.006.
Andren. CM; Wiklund. AKE. (2013). Response of Gammarus pulex and Baetis rhodani to
springtime acid episodes in humic brooks. Sci Total Environ 463-464: 690-699.
http://dx.doi.Org/10.1016/i.scitotenv.2013.06.059.
June 2018
R-7
DRAFT: Do Not Cite or Quote

-------
Angeler. DG; Johnson. RK. (2012). Temporal scales and patterns of invertebrate biodiversity
dynamics in boreal lakes recovering from acidification. Ecol Appl 22: 1172-1186.
http://dx.doi.ore/10.1890/11-1474.1.
Anisfeld. SC: Barnes. RT; Altabet MA: Wu, T. (2007). Isotopic apportionment of atmospheric
and sewage nitrogen sources in two Connecticut rivers. Environ Sci Technol 41: 6363-
6369. http://dx.doi.org/10.1021/esQ70469v.
Anisfeld. SC: Hill. TP. (2012). Fertilization effects on elevation change and belowground carbon
balance in a Long Island Sound tidal marsh. Estuaries Coasts 35: 201-211.
http://dx.doi. org/10.1007/s 1223 7-011 -9440-4.
Appel. KW: Foley. KM: Bash. JO: Pinder. RW: Dennis. RL; Allen. DJ; Pickering. K. (2011). A
multi-resolution assessment of the Community Multiscale Air Quality (CMAQ) model
v4.7 wet deposition estimates for 20022006. Geoscientific Model Development 4: 357-
371. http://dx.doi.org/10.5194/gmd-4-357-2011.
Appel. KW: Napelenok. SL: Foley. KM: Pye. HOT: Hogrefe. C: Luecken. DJ: Bash. JO:
Roselle. SJ: Pleim. JE: Foroutan. H: Hutzell. WT: Pouliot. GA: Sarwar. G: Fahev. KM:
Gantt. B: Gilliam. RC: Heath. NK: Kang. D: Mathur. R: Schwede. DB: Spero. TL:
Wong. DC: Young. JO. (2017). Description and evaluation of the Community Multiscale
Air Quality (CMAQ) modeling system version 5.1. Geoscientific Model Development
10: 1703-1732. http://dx.doi.org/10.5194/gmd-10-1703-2017.
Arens. SJT: Sullivan. PF: Welker. JM. (2008). Nonlinear responses to nitrogen and strong
interactions with nitrogen and phosphorus additions drastically alter the structure and
function of a high arctic ecosystem : Biocomplexity of Artie Tundra Ecosystems. J
Geophys Res 113. http://dx.doi.org/10.1029/2007JG0005Q8.
Arheimer. B: Wittgren. HB. (1994). Modeling the effects of wetlands on regional nitrogen
transport. Ambio23: 378-386.
Armitage. HF: Britton. AJ: van der Wal. R: Pearce. ISK: Thompson. DBA: Woodin. SJ. (2012).
Nitrogen deposition enhances moss growth, but leads to an overall decline in habitat
condition of mountain moss-sedge heath. Global Change Biol 18: 290-300.
http://dx.doi.org/10.1111/i. 1365-2486.2011,02493.x.
Armitage. HF: Britton. AJ: van der Wal. R: Woodin. SJ. (2014). The relative importance of
nitrogen deposition as a driver of Racomitrium heath species composition and richness
across Europe. Biol Conserv 171: 224-231.
http://dx.doi.Org/10.1016/i.biocon.2014.01.039.
Armitage. HF: Britton. AJ: Woodin. SJ: van der Wal. R. (2011). Assessing the recovery potential
of alpine moss-sedge heath: Reciprocal transplants along a nitrogen deposition gradient.
Environ Pollut 159: 140-147. http://dx.doi.Org/10.1016/i.envpol.2010.09.011.
Arneth, A: Harrison, SP: Zaehle, S: Tsigaridis, K: Menon, S: Bartlein, PJ: Feichter, J: Korhola,
A: Kulmala. M: O'Donnell. D: Schurgers. G: Sorvari. S: Vesala. T. (2010). Terrestrial
biogeochemical feedbacks in the climate system. Nat Geosci 3: 525-532.
http://dx.doi.org/10.1038/ngeo9Q5.
June 2018
R-8
DRAFT: Do Not Cite or Quote

-------
Arnett HA; Saros. JE; Mast. MA. (2012). A caveat regarding diatom-inferred nitrogen
concentrations in oligotrophic lakes. J Paleolimnol 47: 277-291.
http://dx.doi.org/10.1007/slQ933-011-9576-z.
Arnold. JR: Hartsell. BE: Luke. WT: Ullah. SMR: Dasgupta. PK: Huev. LG: Tate. P. (2007).
Field test of four methods for gas-phase ambient nitric acid. Atmos Environ 41: 4210-
4226. http://dx.doi.Org/10.1016/i.atmosenv.2006.07.058.
Arora. VK; Boer. GJ: Friedlingstein. P; Ebv. M; Jones. CD: Christian. JR: Bonan. G: Bopp. L;
Brovkin. V: Cadule. P; Haiima. T; Ilyina. T; Lindsay. K; Tjiputra. JF; Wu. T. (2013).
Carbon-concentration and carbon-climate feedbacks in CMIP5 earth system models. J
Clim 26: 5289-5314. http://dx.doi.Org/10.1175/JCLI-D-12-00494.l.
Arroniz-Crespo. M; Leake. JR: Horton. P; Phoenix. GK. (2008). Bryophyte physiological
responses to, and recovery from, long-term nitrogen deposition and phosphorus
fertilisation in acidic grassland. New Phytol 180: 864-874.
http://dx.doi.org/10.1111/i. 1469-8137.2008.02617.x.
Arseneau. KMA; Driscoll. CT; Cummings. CM: Pope. G: Cumming. BF. (2016). Adirondack
(NY, USA) reference lakes show a pronounced shift in chrysophyte species composition
since ca. 1900. J Paleolimnol 56: 349-364. http://dx.doi.org/10.1007/slQ933-016-9922-2.
Ashenden. TW. (1979). The effects of long-term exposures to S02 and N02 pollution on the
growth of Dactylis glomerata L and Poa pratensis L. Environ Pollut 18: 249-258.
http ://dx. doi .org/10.1016/0013 -9327(79N)90020-X.
Asplund, J: Johansson, O; Nybakken, L; Palmqvist, K; Gauslaa, Y. (2010). Simulated nitrogen
deposition influences gastropod grazing in lichens. Ecoscience 17: 83-89.
http://dx.doi.org/10.2980/17-l-3331.
Atalah. J: Crowe. TP. (2012). Nutrient enrichment and variation in community structure on
rocky shores: The potential of molluscan assemblages for biomonitoring. Estuar Coast
Shelf Sci 99: 162-170. http://dx.doi.Org/10.1016/i.ecss.2011.12.034.
Atkeson. T; Greening. Ft; Poor. N. (2007). Bay Region Atmospheric Chemistry Experiment
(BRACE) [Editorial], Atmos Environ 41: 4163-4164.
http://dx.doi.Org/10.1016/i.atmosenv.2006.12.035.
Audet. P; Charest. C. (2007). Heavy metal phytoremediation from a meta-analytical perspective.
Environ Pollut 147: 231-237. http://dx.doi.Org/10.1016/i.envpol.2006.08.011.
Avery. GB. Jr; Kieber. RJ; Witt. M; Willev. JD. (2006). Rainwater monocarboxylic and
dicarboxylic acid concentrations in southeastern North Carolina, USA, as a function of
air-mass back-trajectory. Atmos Environ 40: 16831693.
http://dx.doi.Org/10.1016/i.atmosenv.2005.10.058.
Avis. PG; Mueller. GM; Lussenhop. J. (2008). Ectomycorrhizal fungal communities in two
North American oak forests respond to nitrogen addition. New Phytol 179: 472-483.
http://dx.doi.Org/10.llll/i.1469-8137.2008.02491.x.
June 2018
R-9
DRAFT: Do Not Cite or Quote

-------
Avramescu. ML; Yumvihoze. E; Hintelmann. H; Ridal. J; Fortin. D; Lean. DRS. (2011).
Biogeochemical factors influencing net mercury methylation in contaminated freshwater
sediments from the St. Lawrence River in Cornwall, Ontario, Canada. Sci Total Environ
409: 968-978. http://dx.doi.org/10.1016/i.scitotenv.2010.11.016.
Axler. RP; Reuter. JE. (1996). Nitrate uptake by phytoplankton and periphyton: Whole-lake
enrichments and mesocosm-15N experiments in an oligotrophic lake. Limnology 41:
659-671. http://dx.doi.Org/10.4319/lo.1996.41.4.0659.
Avion DE. (1982). Modeling spore dispersal in a barley crop. Agr Meteorol 26: 215-219.
http://dx.doi.org/10.1016/0002-1571(82^)90032-2.
Azarbad. H; Niklinska. M; van Gestel. CA; van Straalen. NM; Roling. WF; Laskowski. R.
(2013).	Microbial community structure and functioning along metal pollution gradients.
Environ Toxicol Chem 32: 1992-2002. http://dx.doi.org/10.1002/etc.2269.
Babin-Fenske, JJ; Merritt TJS; Gunn, JM; Walsh, T; Lesbarreres, D. (2012). Phylogenetic
analysis of Hyalella colonization in lakes recovering from acidification and metal
contamination. Can J Zool 90: 624-629. http://dx.doi.Org/10.l 139/Z2012-028.
Bachand. PA: Bachand. SM; Fleck. JA; Alpers. CN: Stephenson. M; Windham-Myers. L.
(2014).	Methylmercury production in and export from agricultural wetlands in California,
USA: The need to account for physical transport processes into and out of the root zone.
Sci Total Environ 472: 957-970. http://dx.doi.Org/10.1016/i.scitotenv.2013.l 1.086.
Bachelet. D; Neilson. RP: Lenihan. JM: Drapek. RJ. (2001). Climate change effects on
vegetation distribution and carbon budget in the United States. Ecosystems 4: 164-185.
http ://dx. doi .org/10.1007/s 10021 -001 -0002-7.
Bade. C: Jacob. M; Jungkunst. HF; Leuschner. C: Hauck. M. (2015). Nitrogen mineralization
peaks under closed canopy during the natural forest development cycle of an old-growth
temperate spruce forest. Annals of Forest Science 72: 67-76.
http://dx.doi.org/10.1007/sl3595-014-0394-Q.
Bae. H. -S: Holmes. ME: Chanton. JP; Reddv. KR; Ogram. A. (2015). Distribution, activities,
and interactions of methanogens and sulfate-reducing prokaryotes in the Florida
Everglades. Appl Environ Microbiol 81: 7431-7442.
http://dx.doi.org/10.1128/AEM.01583-15.
Bae. HS: Dierberg. FE; Ogram. A. (2014). Syntrophs dominate sequences associated with the
mercury methylation-related gene hgcA in the water conservation areas of the Florida
Everglades. Appl Environ Microbiol 80: 6517-6526.
http://dx.doi.org/10.1128/AEM.01666-14.
Baez. S: Fargione. J: Moore. DI; Collins. SL; Gosz. JR. (2007). Atmospheric nitrogen deposition
in the northern Chihuahuan desert: Temporal trends and potential consequences. J Arid
Environ 68: 640-651. http://dx.doi.Org/10.1016/i.iaridenv.2006.06.011.
Bahr. A: Ellstrom. M; Akselsson. C: Ekblad. A: Mikusinska. A: Wallander. H. (2013). Growth
of ectomycorrhizal fungal mycelium along a Norway spruce forest nitrogen deposition
gradient and its effect on nitrogen leakage. Soil Biol Biochem 59: 38-48.
http://dx.doi.Org/10.1016/i.soilbio.2013.01.004.
June 2018
R-10
DRAFT: Do Not Cite or Quote

-------
Bahr. A; Ell strom. M; Bergh, J; Wallander. H. (2015). Nitrogen leaching and ectomycorrhizal
nitrogen retention capacity in a Norway spruce forest fertilized with nitrogen and
phosphorus. Plant Soil 390: 323-335. http://dx.doi.org/10.10Q7/sl 1104-015-2408-6.
Bailey. SW: Horslev. SB: Long. RP. (2005). Thirty years of change in forest soils of the
Allegheny Plateau, Pennsylvania. Soil Sci Soc Am J 69: 681-690.
http://dx.doi.org/10.2136/sssai2004.0Q57.
Bailey. SW: Horslev. SB: Long. RP: Hallett. RA. (2004). Influence of edaphic factors on sugar
maple nutrition and health on the Allegheny Plateau. Soil Sci Soc Am J 68: 243-252.
http://dx.doi.org/10.2136/sssai2004.243Q.
Baker. JP: Bernard. DP: Christensen. SW: Sale. MJ: Freda. J: Heltcher. K: Marmorek. D: Rowe.
L: Scanlon. P; Suter. G: Warren-Hicks. WJ: Welbourn. PM. (1990a). Biological effects
of changes in surface water acid-base chemistry. (NAPAP Report 13). Washington, DC:
National Acid Precipitation Assessment Program.
https://ntrl.ntis.gov/NTRL/dashboard/searchResults.xhtml?searchQuery=PB92100494.
Baker. JP: Christensen. SW. (1991). Effects of acidification on biological communities in aquatic
ecosystems. In DF Charles (Ed.), Acidic Deposition and Aquatic Ecosystems (pp. 83-
106). New York, NY: Springer-Verlag.
Baker. JP: Gherini. SA: Christensen. SW: Driscoll. CT: Gallagher. J: Munson. RK: Newton.
RM: Reckhow. KH: Schofield. CL. (1990b). Adirondack lakes survey: An interpretive
analysis of fish communities and water chemistry, 1984-1987. Ray Brook, NY:
Adirondack Lakes Survey Corporation.
http ://www. osti. gov/ scitech/ servlets/purl/6173689.
Baker. JP: Schofield. CL. (1982). Aluminum toxicity to fish in acidic waters. Water Air Soil
Pollut 18: 289-309. http://dx.doi.org/10.1007/BF02419419.
Baker, JP: Schofield, CL. (1985). Acidification impacts on fish populations: A review. In DD
Adams; WP Page (Eds.), Acid Deposition: Environmental, Economic, and Policy Issues
(pp. 183-221). New York, NY: Plenum Press, http://dx.doi.org/10.1007/978-1-4615-
8350-9 12.
Baker, JP: Van Sickle, J: Gagen, CJ: Dewalle, PR: Sharpe, WE; Carline, RF: Baldigo, BP;
Murdoch, PS: Bath, DW: Kretser, WA: Simonin, HA: Wigington, PJ, Jr. (1996). Episodic
acidification of small streams in the northeastern United States: Effects on fish
populations. Ecol Appl 6: 423-437. http://dx.doi.org/10.2307/2269380.
Baker, MA; de Guzman, G; Ostermiller, JD. (2009). Differences in nitrate uptake among benthic
algal assemblages in a mountain stream. J North Am Benthol Soc 28: 24-33.
http://dx.doi.Org/10.1899/07-129.l.
Baker, ME: King, RS. (2010). A new method for detecting and interpreting biodiversity and
ecological community thresholds. Methods Ecol Evol 1: 25-37.
http://dx.doi.org/10.111 l/i.2041-210X.2009.00007.x.
Baldigo, BP: Lawrence, GB. (2001). Effects of stream acidification and habitat on fish
populations of a North American river. Aquat Sci 63: 196-222.
http://dx.doi.org/10.1007/PL000Q1352.
June 2018
R-ll
DRAFT: Do Not Cite or Quote

-------
Baldigo, BP; Lawrence. GB; Bode. RW; Simonin, HA; Roy. KM; Smith. AJ. (2009). Impacts of
acidification on macroinvertebrate communities in streams of the western Adirondack
Mountains, New York, USA. Ecol Indicat 9: 226-239.
http://dx.doi.Org/10.1016/i.ecolind.2008.04.004.
Baldigo. BP; Lawrence. GB; Simonin. HA. (2007). Persistent mortality of brook trout in
episodically acidified streams of the southwestern Adirondack Mountains, New York.
Trans Am Fish Soc 136: 121-134. http://dx.doi.Org/10.1577/T06-043.l.
Baldigo. BP; Murdoch. PS. (1997). Effect of stream acidification and inorganic aluminum on
mortality of brook trout (Salvelinus fontinalis) in the Catskill Mountains, New York. Can
J Fish Aquat Sci 54: 603-615. http://dx.doi.Org/10.l 139/P96-314.
Baldigo. BP; Roy. KM; Driscoll. CT. (2016). Response of fish assemblages to declining acidic
deposition in Adirondack Mountain lakes, 1984-2012. Atmos Environ 146: 223-235.
http://dx.doi.Org/10.1016/i.atmosenv.2016.06.049.
Baldwin. AH. (2013). Nitrogen and phosphorus differentially affect annual and perennial plants
in tidal freshwater and oligohaline wetlands. Estuaries Coasts 36: 547-558.
http://dx.doi.org/10.10Q7/sl2237-011-9445-z.
Ball, BA; Guevara, JA. (2015). The nutrient plasticity of moss-dominated crust in the urbanized
Sonoran Desert. Plant Soil 389: 225-235. http://dx.doi.org/10.1007/slllQ4-014-2355-7.
Ballantyne, AP; Brahnev, J; Fernandez, D; Lawrence, CL; Saros, J; Neff, JC. (2011).
Biogeochemical response of alpine lakes to a recent increase in dust deposition in the
Southwestern, US. Biogeosciences 8: 2689-2706. http://dx.doi.org/10.5194/bg-8-2689-
2011.
Balvanera, P; Pfisterer, AB; Buchmann, N; He, JS; Nakashizuka, T; Raffaelli, D; Schmid, B.
(2006). Quantifying the evidence for biodiversity effects on ecosystem functioning and
services. Ecol Lett 9: 1146-1156. http://dx.doi.org/10.1111/j. 1461-0248.2006.00963,x.
Band, LE; Tague, CL; Groffman, P; Belt, K. (2001). Forest ecosystem processes at the watershed
scale: Hydrological and ecological controls of nitrogen export. Hydrolog Process 15:
2013-2028. http://dx.doi.org/10.1002/hvp.253.
Banzhaf, HS; Burtraw, D; Criscimagna, SC; Cosby, BJ; Evans, DA; Krupnick, AJ; Siikamaeki,
JV. (2016). Policy Analysis: Valuation of Ecosystem Services in the Southern
Appalachian Mountains. Environ Sci Technol 50: 2830-2836.
http://dx.doi.org/10.1021/acs.est.5b03829.
Banzhaf, HS; Burtraw, D; Evans, D; Krupnick, AJ. (2006). Valuation of natural resource
improvements in the Adirondacks. Land Econ 82: 445-464.
http://dx.doi.Org/10.3368/le.82.3.445.
Banzhaf, S; Burtraw, D; Evans, D; Krupnick, AJ. (2004). Valuation of natural resource
improvements in the Adirondacks. Washington, DC: Resources for the Future.
http://portal.nceas.ucsb.edu/working group/science-frameworks-for-ebm/working-group-
meeting-minutes-materials-presentations/september-13-17-2006-meeting/reprints-
exchanged-during-meeting/RFF-RPT-Adirondacks.pdf.
June 2018	R-12	DRAFT: Do Not Cite or Quote

-------
Banzhaf. S; Schaap. M; Kranenburg, R; Manders. AMM; Segers, AJ; Visschedijk. AJH; van Per
Gon. HAC. D; Kuenen. JJP; van Meijgaard. E; van Ulft LH; Cofala. J; Builties. PJH.
(2015). Dynamic model evaluation for secondary inorganic aerosol and its precursors
over Europe between 1990 and 2009. Geoscientific Model Development 8: 1047-1070.
http://dx.doi.org/10.5194/gmd-8-1047-2Q15.
Bao, L; Ma. K; Zhang. S; Lin. L; Ou, L. (2015). Urban dust load impact on gas-exchange
parameters and growth of Sophorajaponica L. seedlings. Plant Soil Environ 61: 309-315.
http://dx.doi.org/10.17221/215/2015-PSE.
Bao. L; Qu. L; Ma. K; Lin. L. (2016). Effects of road dust on the growth characteristics of
Sophorajaponica L. seedlings. J Environ Sci 46: 147-155.
http://dx.doi.Org/10.1016/i.ies.2015.08.031.
Barbier. EB; Hacker. SD; Kennedy. C: Koch. EW: Stier. AC: Silliman. BR. (2011). The value of
estuarine and coastal ecosystem services [Review], Ecol Monogr 81: 169-193.
http://dx.doi.Org/10.1890/10-1510.l.
Bardgett. RD; van Per Putten. W. imH. (2014). Belowground biodiversity and ecosystem
functioning. Nature 515: 505-511. http://dx.doi.org/10.1038/naturel3855.
Bardhan, S; Jose, S; Jenkins, MA; Webster, CR; Udawatta, RP; Stehn, SE. (2012). Microbial
community diversity and composition across a gradient of soil acidity in spruce-fir forests
of the southern Appalachian Mountains. Appl Soil Ecol 61: 60-68.
http://dx.doi.Org/10.1016/i.apsoil.2012.04.010.
Barendregt, A, at: Swarth, CW. (2013). Tidal Freshwater Wetlands: Variation and Changes.
Estuaries Coasts 36: 445-456. http://dx.doi.org/10.1007/sl2237-013-9626-z.
Barker, DH; Vanier, C: Naumburg, E; Charlet, TN: Nielsen, KM: Newingham, BA; Smith, SD.
(2006). Enhanced monsoon precipitation and nitrogen deposition affect leaf traits and
photosynthesis differently in spring and summer in the desert shrub Larrea tridentata.
NewPhytol 169: 799-808. http://dx.doi.org/10.111 l/i.l469-8137.2006.01628.x.
Barker, T; Hatton, K; O'Connor, M; Connor, L; Moss, B. (2008). Effects of nitrate load on
submerged plant biomass and species richness: Results of a mesocosm experiment.
Fundam Appl Limnol 173: 89-100. http://dx.doi.org/10.1127/1863-9135/2008/Q173-
0089.
Barmuta, LA: Cooper, SD: Hamilton, SK; Kratz, KW: Melack, JM. (1990). Responses of
zooplankton and zoobenthos to experimental acidification in a high-elevation lake (Sierra
Nevada, California, USA). Freshw Biol 23: 571-586. http://dx.doi.org/10.1111/j.1365-
2427.1990.tb00296.x.
Barnes, J: Upstill-Goddard, RC. (2011). N20 seasonal distributions and air-sea exchange in UK
estuaries: Implications for the tropospheric N20 source from European coastal waters. J
Geophys Res 116. http://dx.doi.org/10.1029/2009JGQ01156.
Barnthouse, LW: Munns, RM, Jr; Sorensen, MT. (2008). Population-level ecological risk
assessment. Pensacola, FL: Taylor & Francis.
http://www.crcpress.com/product/isbn/9781420Q53326.
June 2018
R-13
DRAFT: Do Not Cite or Quote

-------
Baron. J; Denning. AS. (1993). The influence of mountain meteorology on precipitation
chemistry at low and high elevations of the Colorado Front Range, U.S.A. Atmos
Environ (1967) 27: 2337-2349. http://dx.doi.org/10.1016/0960-1686(93)90402-K.
Baron. JS. (2006). Hindcasting nitrogen deposition to determine ecological critical load. Ecol
Appl 16: 433-439. http://dx.doi.org/10.1890/1051-0761(2006N)016r0433:hndtdal2.0.co:2.
Baron. JS: Driscoll. CT; Stoddard. JL. (201 la). Inland surface water. In Assessment of Nitrogen
Deposition Effects and Empirical Critical Loads of Nitrogen for Ecoregions of the United
States (pp. 209-228). (NRS-80). Newtown Square, PA: U.S. Department of Agriculture,
Forest Service, Northern Research Station.
http://www.nrs.fs.fed.us/pubs/gtr/gtr nrs80.pdf?
Baron. JS: Driscoll. CT: Stoddard. JL: Richer. EE. (201 lb). Empirical critical loads of
atmospheric nitrogen deposition for nutrient enrichment and acidification of sensitive US
lakes. Bioscience 61: 602-613. http://dx.doi.org/10.1525/bio.2011.61.8.6.
Baron. JS: Hall. EK; Nolan. BT; Finlav. JC: Bernhardt. ES: Harrison. JA; Chan. F; Bover. EW.
(2012). The interactive effects of excess reactive nitrogen and climate change on aquatic
ecosystems and water resources of the United States. Biogeochemistry 114: 71-92.
http://dx.doi.org/10.1007/slQ533-012-9788-v.
Baron. JS: Oiima. DS: Holland. EA; Parton. WJ. (1994). Analysis of nitrogen saturation
potential in Rocky Mountain tundra and forest: Implications for aquatic systems.
Biogeochemistry 27: 61-82. http://dx.doi.org/10.1007/BF000Q2571.
Baron, JS: Rueth, HM; Wolfe, AM; Nydick, KR; Allstott, EJ; Minear, JT; Moraska, B. (2000).
Ecosystem responses to nitrogen deposition in the Colorado Front Range. Ecosystems 3:
352-368. http://dx.doi.org/10.1007/sl00210000Q32.
Baron, JS: Schmidt, TM: Hartman, MP. (2009). Climate-induced changes in high elevation
stream nitrate dynamics. Global Change Biol 15: 1777-1789.
http://dx.doi.org/10.111 l/i.l365-2486.2009.01847.x.
Bartell, SM. (2007). Ecosystem effects modeling. In GW Suterll (Ed.), Ecological risk
assessment (2nd ed., pp. 413-432). Boca Raton, FL: CRC Press.
Bartlett, KB: Harriss, RC: Sebacher, DI. (1985). Methane flux from coastal salt marshes. J
Geophys Res Atmos 90: 5710-5720. http://dx.doi.org/10.1029/JD090iD03pQ5710.
Barton, A: Hales, B: Waldbusser, GG: Langdon, C; Feely, RA. (2012). The Pacific oyster,
Crassostrea gigas, shows negative correlation to naturally elevated carbon dioxide levels:
Implications for near-term ocean acidification effects. Limnol Oceanogr 57: 698-710.
http://dx.doi.Org/10.4319/lo.2012.57.3.0698.
Barton, A: Waldbusser, GG: Feely, RA: Weisberg, SB: Newton, JA: Hales, B: Cudd, S:
Eudeline, B: Langdon, CJ: Jefferds, I: King, T; Suhrbier, A: McLaughlin, K. (2015).
Impacts of coastal acidification on the Pacific Northwest shellfish industry and adaptation
strategies implemented in response. Oceanography 28: 146-159.
http://dx.doi.org/10.5670/oceanog.2015.38.
June 2018
R-14
DRAFT: Do Not Cite or Quote

-------
Bash. JO; Cooter, EJ; Dennis. RL; Walker. JT; Pleim, JE. (2012). Evaluation of a regional air-
quality model with bi-directional NH3 exchange coupled to an agro-ecosystem model.
Biogeosciences Discussions 9: 11375-11401. http://dx.doi.org/10.5194/bgd-9-11375-
2012.
Bash. JO; Cooter. EJ; Dennis. RL; Walker. JT; Pleim. JE. (2013). Evaluation of a regional air-
quality model with bidirectional NH3 exchange coupled to an agroecosystem model.
Biogeosciences 10: 1635-1645. http://dx.doi.org/10.5194/bg-10-1635-2013.
Bassin. S; Kach. D; Valsangiacomo. A; Mayer. J; Oberholzer. H. -R; Yolk. M; Fuhrer. J. (2015).
Elevated ozone and nitrogen deposition affect nitrogen pools of subalpine grassland.
EnvironPollut 201: 67-74. http://dx.doi.Org/10.1016/i.envpol.2015.02.038.
Bassin. S; Schalaida. J; Vogel. A; Suter. M. (2012). Different types of sub-alpine grassland
respond similarly to elevated nitrogen deposition in terms of productivity and sedge
abundance. J Veg Sci 23: 1024-1034. http://dx.doi.org/10.Ill 1/i.1654-
1103.2012.01422.x.
Bassin. S; Yolk. M; Fuhrer. J. (2013). Species composition of subalpine grassland is sensitive to
nitrogen deposition, but not to ozone, after seven years of treatment. Ecosystems 16:
1105-1117. http://dx.doi.org/10.1007/sl0021-013-9670-3.
Bassin. S; Yolk. M; Suter. M; Buchmann. N; Fuhrer. J. (2007). Nitrogen deposition but not
ozone affects productivity and community composition of subalpine grassland after 3 yr
of treatment. New Phytol 175: 523-534. http://dx.doi.Org/10.l 111/i. 1469-
8137.2007.02140.x.
Bassin. S; Werner. RA; Sorgel. K; Yolk. M; Buchmann. N; Fuhrer. J. (2009). Effects of
combined ozone and nitrogen deposition on the in situ properties of eleven key plant
species of a subalpine pasture. Oecologia 158: 747-756.
http://dx.doi.org/10.1007/s00442-0Q8-1191-v.
Bassirirad. H; Lussenhop. JF; Sehtiva. HL; Borden. KK. (2015). Nitrogen deposition potentially
contributes to oak regeneration failure in the Midwestern temperate forests of the USA.
Oecologia 177: 53-63. http://dx.doi.org/10.1007/s00442-014-3119-z.
Basto, S; Thompson, K; Phoenix, G; Sloan, V; Leake, J; Rees, M. (2015a). Long-term nitrogen
deposition depletes grassland seed banks. Nat Commun 6: 6185.
http://dx.doi.org/10.1038/ncomms7185.
Basto, S; Thompson, K; Rees, M. (2015b). The effect of soil pH on persistence of seeds of
grassland species in soil. Plant Ecol 216: 1163-1175. http://dx.doi.org/10.10Q7/sll258-
015-0499-z.
Bates, AL; Orem, WH; Harvey, JW; Spiker, EC. (2002). Tracing sources of sulfur in the Florida
Everglades. J Environ Qual 31: 287-299. http://dx.doi.org/10.2134/ieq2002.2870.
Bates, NR; Peters, AJ. (2007). The contribution of atmospheric acid deposition to ocean
acidification in the subtropical North Atlantic Ocean. Mar Chem 107: 547-558.
http://dx.doi.Org/10.1016/i.marchem.2007.08.002.
June 2018
R-15
DRAFT: Do Not Cite or Quote

-------
Battarbee. RW; Shi 11 and, EM; Kern an. M; Monteith. DT; Curtis. CJ. (2014). Recovery of
acidified surface waters from acidification in the United Kingdom after twenty years of
chemical and biological monitoring (1988-2008). Ecol Indicat 37: 267-273.
http://dx.doi.Org/10.1016/i.ecolind.2013.10.011.
Battles. JJ; Fahev, TJ; Driscoll, CT. Jr; Blum. JD; Johnson. CE. (2014). Restoring soil calcium
reverses forest decline. Environ Sci Technol 1: 15-19.
http://dx.doi.ore/10.1021/ez400033d.
Bauer. GA; Bazzaz. FA: Minocha. R; Long. S: Magill. A: Aber. J: Berntson. GM. (2004).
Effects of chronic N additions on tissue chemistry, photosynthetic capacity, and carbon
sequestration potential of a red pine (Pinus resinosa Ait) stand in the NE United States.
For Ecol Manage 196: 173-186. http://dx.doi.Org/10.1016/i.foreco.2004.03.032.
Baum. MM: Kiyomiya. ES: Kumar. S: Lappas. AM: Kapinus. VA; Lord. HC. (2001).
Multicomponent remote sensing of vehicle exhaust by dispersive absorption
spectroscopy. 2. Direct on-road ammonia measurements. Environ Sci Technol 35: 3735-
3741.
Baumann. K; Marschner. P; Smernik. RJ; Baldock. JA. (2009). Residue chemistry and microbial
community structure during decomposition of eucalypt, wheat and vetch residues. Soil
Biol Biochem 41: 1966-1975. http://dx.doi.Org/10.1016/i.soilbio.2009.06.022.
Baumgardner; Lavery. TF; Rogers. CM: Isil. SS. (2002). Estimates of the atmospheric deposition
of sulfur and nitrogen species:Clean air status and trends network, 19902000. Environ Sci
Technol 36: 2614-2629. http://dx.doi.org/10.1021/es011146g.
Baustian. MM: Rabalais. NN. (2009). Seasonal composition of benthic macroinfauna exposed to
hypoxia in the Northern Gulf of Mexico. Estuaries Coasts 32: 975-983.
http://dx.doi.org/10.1007/sl2237-0Q9-9187-3.
Bayley, SE; Schindler, DW: Parker, BR; Stainton, MP; Beaty, KG. (1992). Effects of forest fire
and drought on acidity of a base-poor boreal forest stream: Similarities between climatic
warming and acidic precipitation. Biogeochemistry 17: 191-204.
http://dx.doi. org/10.1007/BF00004041.
Beaulieu, JJ: Tank, JL; Hamilton, SK; Wollheim, WM; Hall, RO, Jr; Mulholland, PJ; Peterson,
BJ; Ashkenas, LR; Cooper, LW; Dahm, CN; Dodds, WK; Grimm, NB; Johnson, SL;
McDowell, WH: Poole, GC; Valett, HM; Arango, CP; Bernot, MJ; Burgin, AJ;
Crenshaw, CL; Helton, AM; Johnson, LT; O'Brien, JM; Potter, JD; Sheiblev, RW;
Sobota, DJ; Thomas, SM. (2011). Nitrous oxide emission from denitrification in stream
and river networks. Proc Natl Acad Sci USA 108: 214-219.
http://dx.doi.org/10.1073/pnas.1011464108.
Beaver, JR; Manis, EE; Loftin, KA; Graham, JL; Pollard, AI; Mitchell, RM. (2014). Land use
patterns, ecoregion, and microcystin relationships in US lakes and reservoirs: A
preliminary evaluation. Harmful Algae 36: 57-62.
http://dx.doi.Org/10.1016/i.hal.2014.03.005.
Bebber, DP; Watkinson, SC; Boddv, L; Darrah, PR. (2011). Simulated nitrogen deposition
affects wood decomposition by cord-forming fungi. Oecologia 167: 1177-1184.
http://dx.doi.org/10.1007/s00442-011-2Q57-2.
June 2018
R-16
DRAFT: Do Not Cite or Quote

-------
Beck. WS; Rugenski, AT; Poff. NL. (2017). Influence of experimental, environmental, and
geographic factors on nutrient-diffusing substrate experiments in running waters. Freshw
Biol 62: 1667-1680. http://dx.doi.org/10.1111/fwb. 12989.
Becker. JC: Groeger. AW: Nowlin. WH; Chumchal. MM: Hahn. D. (2011). Spatial variability in
the speciation and bioaccumulation of mercury in an arid subtropical reservoir ecosystem.
Environ Toxicol Chem 30: 2300-2311. http://dx.doi.org/10.1002/etc.626.
Bedison. JE; McNeil. BE. (2009). Is the growth of temperate forest trees enhanced along an
ambient nitrogen deposition gradient? Ecology 90: 1736-1742.
http://dx.doi.Org/10.1890/08-0792.l.
Beebv. A: Richmond. L. (2002). Evaluating Helix aspersa as a sentinel for mapping metal
pollution. Ecol Indicat 1: 261-270. http://dx.doi.org/10.1016/S1470-160X(02N)00022-5.
Beem. KB: Raia. S: Schwandner. FM; Taylor. C: Lee. T; Sullivan. AP; Carrico. CM:
McMeeking, GR; Day, D; Levin, E; Hand, J: Kreidenweis, SM; Schichtel, B; Malm, WC;
Collett, J, r. (2010). Deposition of reactive nitrogen during the Rocky Mountain Airborne
Nitrogen and Sulfur (RoMANS) study. Environ Pollut 158: 862-872.
http://dx.doi.Org/10.1016/i.envpol.2009.09.023.
Beem an, P. (2013). Nitrate spike tests Des Moines water supplies.
Beer, R; Shephard, MW; Kulawik, SS: Clough, SA; Eldering, A: Bowman, KW: Sander, SP;
Fisher, BM; Payne, VH; Luo, M; Osterman, GN: Worden, JR. (2008). First satellite
observations of lower tropospheric ammonia and methanol. Geophys Res Lett 35:
L09801. http://dx.doi.org/10.1029/2008GLQ33642.
Beggs, GL; Gunn, JM. (1986). Response of lake trout (Salvelinus namaycush) and brook trout (S
fontinalis) to surface water acidification in Ontario. Water Air Soil Pollut 30: 711-717.
http://dx.doi.org/10.1007/BF003Q3336.
Beidleman, LH; Beidleman, RG: Willard, BE: Nelson, RA. (2000). Plants of Rocky Mountain
National Park: A complete revision of Ruth Ashton Nelson's popular manual. Hlenea,
MT: Rocky Mountain Nature Association & Falcon Pub.
Beier, C: Blanck, K; Bredemeier, M; Lamersdorf, N: Rasmussen, L; Xu, Y, -J. (1998). Field-
scale 'clean rain' treatments to two Norway spruce stands within the EXMAN project:
Effects on soil solution chemistry, foliar nutrition and tree growth. For Ecol Manage 101:
111-123. http://dx.doi.org/10.1016/S0378-1127(97^)00129-1.
Beier, CM: Caputo, J: Lawrence, GB; Sullivan, TJ. (2017). Loss of ecosystem services due to
chronic pollution of forests and surface waters in the Adirondack region (USA). J
Environ Manage 191: 19-27. http://dx.doi.Org/10.1016/i.ienvman.2016.12.069.
Beier, CM; Woods, AM; Hotopp, KP; Gibbs, JP; Mitchell, MJ; Dovciak, M; Leopold, DJ;
Lawrence, GB: Page, BP. (2012). Changes in faunal and vegetation communities along a
soil calcium gradient in northern hardwood forests. Can J For Res 42: 1141-1152.
http://dx.doi.org/10.1139/X2012-071.
June 2018
R-17
DRAFT: Do Not Cite or Quote

-------
Bell MP; Phelan. J; Blett TF; Landers. D; Nahlik. AM; Van Houtven. G; Davis. C; Clark. CM;
Hewitt. J. (2017). A framework to quantify the strength of ecological links between an
environmental stressor and final ecosystem services. Ecosphere 8.
http://dx.doi. org/10.1002/ecs2.1806.
Bellinger. BJ; Jicha. TM; Lehto. LP; Seifert-Monson. LR; Bolgrien, DW; Starry. MA; Angradi,
TR; Pearson. MS; Elonen. C; Hill. BH. (2014). Sediment nitrification and denitrification
in a Lake Superior estuary. J Great Lakes Res 40: 392-403.
http://dx.doi.Org/10.1016/i.iglr.2014.03.012.
Belnap. J; Phillips. SL; Flint. S; Money. J; Caldwell. M. (2008). Global change and biological
soil crusts: Effects of ultraviolet augmentation under altered precipitation regimes and
nitrogen additions. Global Change Biol 14: 670-686. http://dx.doi.Org/10.l 11 l/j.1365-
2486.2007.01509.x.
Belnap. J; Sigal. L; Moir. W; Eversman. S. (1993). Identification of sensitive species, Lichens as
bioindicators of air quality. In LS Huckaby (Ed.), Lichens as Bioindicators of Air Quality
(general technical report RM-224 ed., pp. 67-88). Fort Collins, CO: U.S. Department of
Agriculture, Forest Service, Rocky Mountain Forest and Range Experimental Station.
Belvazid. S; Kurz. D; Braun. S; Sverdrup. H; Rihm. B; Hettelingh. J. -P. (201 la). A dynamic
modelling approach for estimating critical loads of nitrogen based on plant community
changes under a changing climate. Environ Pollut 159: 789-801.
http://dx.doi.Org/10.1016/i.envpol.2010.l 1.005.
Belvazid. S; Sverdrup. H; Kurz. D; Braun. S. (201 lb). Exploring ground vegetation change for
different deposition scenarios and methods for estimating critical loads for biodiversity
using the ForSAFE-VEG model in Switzerland and Sweden. Water Air Soil Pollut 216:
289-317. http://dx.doi.org/10.1007/sll270-010-Q534-6.
Bender. J; Wei gel. H. -J. (2011). Changes in atmospheric chemistry and crop health. In
Sustainable Agriculture Volume 2. Netherlands: Springer Netherlands.
http://dx.doi.org/10.1007/978-94-007-Q394-0 22.
Benedict. KB; Carrico. CM; Kreidenweis. SM; Schichtel. B; Malm. WC; Collett. JL. Jr. (2013).
A seasonal nitrogen deposition budget for Rocky Mountain National Park. Ecol Appl 23:
1156-1169. http://dx.doi.Org/10.1890/12-1624.l.
Bennett. DA; Goble. RL; Linthurst. RA. (1985). The acidic deposition phenomenon and its
effects: critical assessment document. Bennett, DA; Goble, RL; Linthurst, RA.
Benoit, JM; Gilmour, CC; Heyes, A; Mason, RP; Miller, CL. (2003). Geochemical and
biological controls over methylmercury production and degradation in aquatic
ecosystems. In Y Chai; OC Braids (Eds.), Biogeochemistry of Environmentally
Important Trace Elements (ACS Symposium Series no. 835 ed., pp. 262-297).
Washington, D.C.: American Chemical Society. http://dx.doi.org/10.1021/bk-20Q3-
Q835.ch019.
Benoit, JM; Gilmour, CC; Mason, RP; Riedel, GS; Riedel, GF. (1998). Behavior of mercury in
the Patuxent River estuary. Biogeochemistry 40: 249-265.
http://dx.doi.Org/10.1023/A:1005905700864.
June 2018
R-18
DRAFT: Do Not Cite or Quote

-------
Benson. JL; Schlezinger, D: Howes. BL. (2013). Relationship between nitrogen concentration,
light, and Zostera marina habitat quality and survival in southeastern Massachusetts
estuaries. J Environ Manage 131: 129-137.
http://dx.doi.Org/10.1016/i.ienvman.2013.09.033.
Bentz. BJ: Regniere. J: Fettig. CJ: Hansen. EM: Haves. JL: Hicke. JA; Kelsev. RG: Negron. JF;
Sevbold. SJ. (2010). Climate Change and Bark Beetles of the Western United States and
Canada: Direct and Indirect Effects. Bioscience 60: 602-613.
http://dx.doi.Org/10.1525/bio.2010.60.8.6.
Berdieb. L; Pollet. T; Chardon. C: Jacquet. S. (2013). Spatio-temporal changes in the structure of
archaeal communities in two deep freshwater lakes. FEMS Microbiol Ecol 86: 215-230.
http://dx.doi.org/10.1111/1574-6941.12154.
Berg. GM; Driscoll. S: Havashi. K; Ross. M; Kudela. R. (2017). Variation in growth rate, carbon
assimilation, and photosynthetic efficiency in response to nitrogen source and
concentration in phytoplankton isolated from upper San Francisco Bay. J Phycol 53: 664-
679. http://dx.doi.org/10.Ill 1/jpy. 12535.
Bergametti. G: Remoudaki. E; Losno. R; Steiner. E; Chatenet. B; Buat-Menard. P. (1992).
Source, transport and deposition of atmospheric phosphorus over the northwestern
Mediterranean. J Atmos Chem 14: 501-513. http://dx.doi.org/10.1007/BFQ0115254.
Berger. TW: Turtscher. S: Berger. P; Lindebner. L. (2016). A slight recovery of soils from Acid
Rain over the last three decades is not reflected in the macro nutrition of beech (Fagus
sylvatica) at 97 forest stands of the Vienna Woods. Environ Pollut 216: 624-635.
http://dx.doi.Org/10.1016/i.envpol.2016.06.024.
Berglund. AM; Rainio. MJ; Eeva. T. (2015). Temporal trends in metal pollution: using bird
excrement as indicator. PLoS ONE 10: eOl 17071.
http://dx.doi.org/10.1371/iournal.pone.0117Q71.
Berglund. AMM; Koivula. MJ: Eeva. T. (2011). Species- and age-related variation in metal
exposure and accumulation of two passerine bird species. Environ Pollut 159: 2368-2374.
http://dx.doi.Org/10.1016/i.envpol.2011.07.001.
Berglund, AMM; Nyholm, NEI. (2011). Slow improvements of metal exposure, health- and
breeding conditions of pied flycatchers (Ficedula hypoleuca) after decreased industrial
heavy metal emissions. Sci Total Environ 409: 4326-4334.
http://dx.doi. org/10.1016/i. scitotenv.2011.07.004.
Berglund, AMM; Rainio, MJ: Eeva, T. (2012). Decreased metal accumulation in passerines as a
result of reduced emissions. Environ Toxicol Chem 31: 1317-1323.
http://dx.doi.org/10.1002/etc.1814.
Bergman, HL; Mattice, JS: Brown, DJA. (1988). Lake acidification and fisheries project: Adult
brook trout (Salvelinus fontinalis). Can J Fish Aquat Sci 45: 1561-1562.
http://dx.doi.Org/10.l 139/f88-185.
Bergman, I; Bishop, K; Tu, Q: Freeh, W: Akerblom, S: Nilsson, M. (2012). The influence of
sulphate deposition on the seasonal variation of peat pore water methyl Hg in a Boreal
Mire. PLoS ONE 7: e45547. http://dx.doi.org/10.1371/iournal.pone.0Q45547.
June 2018
R-19
DRAFT: Do Not Cite or Quote

-------
Bergstrom. A. -K. (2010). The use of TN:TP and DIN:TP ratios as indicators for phytoplankton
nutrient limitation in oligotrophic lakes affected by N deposition. Aquat Sci 72: 277-281.
http://dx.d0i.0rg/l 0.1007/s00027-010-0132-0.
Bergstrom. A. -K; Blomqvist. P; Jansson. M. (2005). Effects of atmospheric nitrogen deposition
on nutrient limitation and phytoplankton biomass in unproductive Swedish lakes.
Limnology 50: 987-994. http://dx.doi.Org/10.4319/lo.2005.50.3.0987.
Bergstrom. A. -K; Jansson. M. (2006). Atmospheric nitrogen deposition has caused nitrogen
enrichment and eutrophication of lakes in the northern hemisphere. Global Change Biol
12: 635-643. http://dx.doi.Org/10.l 111/i. 1365-2486.2006.01129.x.
Bergstrom. A. -K; Jonsson. A: Jansson. M. (2008). Phytoplankton responses to nitrogen and
phosphorus enrichment in unproductive Swedish lakes along a gradient of atmospheric
nitrogen deposition. Aquatic Biology 4: 55-64. http://dx.doi.org/10.3354/ab00099.
Bergstrom, A, nnK; Karlsson, D; Karlsson, J, an; Vrede, T. (2015). N-limited consumer growth
and low nutrient regeneration N:P ratios in lakes with low N deposition. Ecosphere 6.
http://dx.doi.Org/10.1890/ES14-00333.l.
Bergstrom. AK; Faithfull. C: Karlsson. D; Karlsson. J. (2013). Nitrogen deposition and warming:
Effects on phytoplankton nutrient limitation in subarctic lakes. Global Change Biol 19:
2557-2568. http://dx.doi.org/10.1111/gcb. 12234.
Bernard. RJ; Mortazavi. B; Kleinhuizen. AA. (2015). Dissimilatory nitrate reduction to
ammonium (DNRA) seasonally dominates N03 (-) reduction pathways in an
anthropogenically impacted sub-tropical coastal lagoon. Biogeochemistry 125: 47-64.
http://dx.doi.org/10.1007/slQ533-015-0111-6.
Berndt. T; Jokinen. T; Mauldin. R. Ill: Petaia. T; Herrmann. H; Junninen. H; Paasonen. P;
Worsnop. PR: Sipila. M. (2012). Gas-phase ozonolysis of selected olefins: The yield of
stabilized criegee intermediate and the reactivity toward S02. J Phys Chem Lett 3: 2892-
2896. http://dx.doi.org/10.1021/iz301158u.
Berthelsen. BO: Olsen. RA; Steinnes. E. (1995). Ectomycorrhizal heavy metal accumulation as a
contributing factor to heavy metal levels in organic surface soils. Sci Total Environ 170:
141-149. http://dx.doi.org/10.1016/0048-9697(95)04701-2.
Beswick. KM: Hargreaves. KJ; Gallagher. MW; Choularton. TW: Fowler. D. (1991). Size-
resolved measurements of cloud droplet deposition velocity to a forest canopy using an
eddy correlation technique. Q J Roy Meteorol Soc 117: 623-645.
http://dx.doi.org/10.1002/qi.49711749910.
Bethers. S: Day. ME: Wiersma. GB; Fernandez. IJ; Elvir. JA. (2009). Effects of chronically
elevated nitrogen and sulfur deposition on sugar maple saplings: Nutrition, growth and
physiology. For Ecol Manage 258: 895-902.
http://dx.doi.Org/10.1016/i.foreco.2009.03.024.
Bettez. ND; Marino. R; Howarth. RW: Davidson. EA. (2013). Roads as nitrogen deposition hot
spots. Biogeochemistry 114: 149-163. http://dx.doi.org/10.1007/slQ533-013-9847-z.
June 2018
R-20
DRAFT: Do Not Cite or Quote

-------
Beyer. WN; Green. CE; Beyer. M; Chanev. RL. (2013). Phytotoxicity of zinc and manganese to
seedlings grown in soil contaminated by zinc smelting. Environ Pollut 179: 167-176.
http://dx.doi.ore/10.1016/j.envpol. 2013.04.013.
Beyer. WN: Krafft. C: Klassen. S: Green. CE: Chanev. RL. (2011). Relating Injury to the Forest
Ecosystem Near Palmerton, PA, to Zinc Contamination From Smelting. Arch Environ
Contam Toxicol 61: 376-388. http://dx.doi.org/10.1007/s00244-010-9640-0.
Bigham. GN: Murray. KJ: Masue-Slowev. Y; Henry. EA. (2017). Biogeochemical controls on
methylmercury in soils and sediments: Implications for site management [Review], Integr
Environ Assess Manag 13: 249-263. http ://dx. doi. org/10.1002/ieam. 1822.
Billen. G: Gamier. J: Lassaletta. L. (2013). The nitrogen cascade from agricultural soils to the
sea: Modelling nitrogen transfers at regional watershed and global scales. Philos Trans R
Soc Lond B Biol Sci 368: 20130123. http://dx.doi.org/10.1098/rstb.2013.0123.
Billett, MF; Parker-Jervis, F; Fitzpatrick, EA; Cresser, MS. (1990). Forest soil chemical changes
between 1949/50 and 1987. Soil Sci Soc Am J 41: 133-145.
http://dx.doi.Org/10.llll/i.1365-2389.1990.tb00051.x.
Bilodeau-Gauthier. S: Houle. D; Gagnon. C: Cote. B; Messier. C. (2011). Assessment of sugar
maple tree growth in relation to the partitioning of elements in xylem along a soil acidity
gradient. For Ecol Manage 261: 95-104. http://dx.doi.Org/10.1016/i.foreco.2010.09.035.
Birch. MBL; Gramig. BM; Moomaw. WR: Doering. OC. Ill: Reeling. CJ. (2011). Why metrics
matter: Evaluating policy choices for reactive nitrogen in the Chesapeake Bay Watershed.
Environ Sci Technol 45: 168-174. http://dx.doi.org/10.1021/esl01472z.
Bishop. JG: O'Hara. NB; Titus. JH; Apple. JL; Gill. RA; Wynn. L. (2010). N-P co-limitation of
primary production and response of arthropods to N and P in early primary succession on
Mount St. Helens volcano. PLoS ONE 5: el3598.
http://dx.doi.org/10.1371/iournal.pone.0013598.
Bishop. K; Rapp. L; Kohler. S: Korsman. T. (2008). Testing the steady-state water chemistry
model predictions of pre-industrial lake pH with paleolimnolognical data from northern
Sweden. Sci Total Environ 407: 723-729.
http://dx.doi.Org/10.1016/i.scitotenv.2008.10.007.
Bladon. KD; Emelko. MB: Silins. U: Stone. M. (2014). Wildfire and the Future of Water Supply.
Environ Sci Technol 48: 8936-8943. http://dx.doi.org/10.1021/es50Q130g.
Blake. L; Goulding. KWT; Mott. CJB; Johnston. AE. (1999). Changes in soil chemistry
accompanying acidification over more than 100 years under woodland and grass at
Rothamsted Experimental Station, UK. Eur J Soil Sci 50: 401-412.
http://dx.doi.Org/10.1046/i.1365-2389.1999.00253.x.
Blanchard. CL; Hidv. GM; Tanenbaum. S: Edgerton. ES: Hartsell. BE. (2013). The Southeastern
Aerosol Research and Characterization (SEARCH) study: Temporal trends in gas and
PM concentrations and composition, 1999-2010. J Air Waste Manag Assoc 63: 247-259.
http://dx.doi.org/10.1080/10962247.2Q12.748523.
June 2018
R-21
DRAFT: Do Not Cite or Quote

-------
Blanke. V; Bassin. S; Yolk. M; Fuhrer. J. (2012). Nitrogen deposition effects on subalpine
grassland: The role of nutrient limitations and changes in mycorrhizal abundance. Acta
Oecol 45: 57-65. http://dx.doi.ore/10.1016/j.actao.2012.09.002.
Blaustein. AR; Wake. DB. (1990). Declining amphibian populations: A global phenomenon?
Trends Ecol Evol 5: 203-204. http://dx.doi.org/10.1016/0169-5347(90N)90129-2.
Blett TF; Lynch. JA; Pardo. LH; Huber. C: Haeuber. R; Pouvat R. (2014). FOCUS: A pilot
study for national-scale critical loads development in the United States. Environ Sci Pol
38: 225-236. http://dx.doi.Org/10.1016/i.envsci.2013.12.005.
Bloch, I; Weyhenmeyer, GA. (2012). Long-term changes in physical and chemical conditions of
nutrient-poor lakes along a latitudinal gradient: Is there a coherent phytoplankton
community response? Aquat Sci 74: 77-85. http://dx.doi.org/10.1007/s00027-011-0197-
4.
Blomqvist, P; Pettersson, A; Hyenstrand, P. (1994). Ammonium-nitrogen: A key regulatory
factor causing dominance of non-nitrogen-fixing cyanobacteria in aquatic systems. Arch
Hydrobiol 132: 141-164.
Bloom. AJ: Chapin. FS. Ill: Moonev. HA. (1985). Resource limitation in plants: An economic
analogy. Annu Rev Ecol Evol Systemat 16: 363-392.
http://dx.doi.Org/10.l 146/annurev.es. 16.110185.002051.
Bloom. NS: Watras. CJ: Hurley. JP. (1991). Impact of acidification on the methylmercury cycle
of remote seepage lakes. Water Air Soil Pollut 56: 477-491.
http://dx.doi.org/10.1007/BF0Q342293.
Bobbink. R; Ashmore. M; Braun. S: Fltickiger. W: Van den Wyngaert. IJJ. (2003). Empirical
nitrogen critical loads for natural and semi-natural ecosystems: 2002 update. In B
Achermann; R Bobbink (Eds.), Empirical Critical Loads for Nitrogen. Berne,
Switzerland: Swiss Agency for Environment, Forest and Landscape SAEFL.
Bobbink. R; Boxman. D; Fremstad. E; Heil. G: Houdijk. A: Roelofs. J. (1992). Critical loads for
nitrogen eutrophication of terrestrial and wetland ecosystems based upon changes in
vegetation and fauna. In P Grennfelt; E Thornelof (Eds.), Critical Loads for Nitrogen
(Nord 1992:41 ed., pp. 111-159). Copenhagen, Denmark: Nordic Council of Ministers.
Bobbink. R; Hicks. K; Galloway. J: Spranger. T; Alkemade. R; Ashmore. M; Bustamante. M;
Cinderbv. S: Davidson. E; Dentener. F; Emett. B; Erisman. JW: Fenn. M; Gilliam. F;
Nordin. A: Pardo. L; W. DV. (2010). Global assessment of nitrogen deposition effects on
terrestrial plant diversity: A synthesis. Ecol Appl 20: 30-59. http://dx.doi.org/10.1890/08-
1140.1.
Bobbink. R; Hornung. M; Roelofs. JGM. (1998). The effects of air-borne nitrogen pollutants on
species diversity in natural and semi-natural European vegetation. J Ecol 86: 717-738.
http://dx.doi.Org/10.1046/i.1365-2745.1998.8650717.x.
Boersma. KF; Eskes. HJ; Brinksma. EJ. (2004). Error analysis for tropospheric N02 retrieval
from space. J Geophys Res 109: D04311. http://dx.doi.org/10.1029/2003JD0Q3962.
June 2018
R-22
DRAFT: Do Not Cite or Quote

-------
Boesch. DF; Coles. VJ; Kimmel. DG; Miller. WD. (2007). V. Ramifications of climate change
for Chesapeake Bay hypoxia. In Regional Impacts of Climate Change: Four Case Studies
in the United States (pp. 57-70). Arlington, VA: Pew Center on Global Climate Change.
Bogdziewicz. M; Crone. EE; Steele. MA; Zwolak. R. (2017). Effects of nitrogen deposition on
reproduction in a masting tree: benefits of higher seed production are trumped by
negative biotic interactions. J Ecol 105: 310-320. http://dx.doi.Org/10.l 111/1365-
2745.12673.
Bohlen. PJ; Groffman. PM; Fahev. TJ; Fisk. MC; Suarez. E; Pelletier. DM; Fahev. RT. (2004).
Ecosystem consequences of exotic earthworm invasion of north temperate forests.
Ecosystems 7: 1-12. http://dx.doi.org/10.1007/sl0021-003-Q126-z.
Boiano. DM; Weeks. DP; Hemry. T. (2005). Sequoia and Kings Canyon National Parks,
California: Water resources information and issues overview report. (Tech. Report
NPS/NRWRD/NRTR-2005/333). Washington, DC: National Park Service, U.S.
Department of the Interior.
https://www.nature.nps.gov/water/planning/Info Issuesoverview reports/seki wriio fina
1 High.pdf.
Boiarczuk. K; Kieliszewska-Rokicka. B. (2010). Effect of ectomycorrhiza on Cu and Pb
accumulation in leaves and roots of silver birch (Betula pendula Roth.) seedlings grown
in metal-contaminated soil. Water Air Soil Pollut 207: 227-240.
http://dx.doi.org/10.1007/sll270-009-Q131-8.
Bollmann. A; Bulleriahn. GS; Mckav. RM. (2014). Abundance and Diversity of Ammonia-
Oxidizing Archaea and Bacteria in Sediments of Trophic End Members of the Laurentian
Great Lakes, Erie and Superior. PLoS ONE 9: e97068.
http://dx.doi.org/10.1371/iournal.pone.0097Q68.
Boltersdorf. SH; Pesch. R; Werner. W. (2014). Comparative use of lichens, mosses and tree bark
to evaluate nitrogen deposition in Germany. Environ Pollut 189: 43-53.
http://dx.doi.Org/10.1016/i.envpol.2014.02.017.
Bonan. GB; Levis. S. (2010). Quantifying carbon-nitrogen feedbacks in the Community Land
Model (CLM4). Geophys Res Lett 37: L07401.
http://dx.doi.org/10.1029/2010GL04243Q.
Bond-Lambertv. B; Thomson. A. (2010). Temperature-associated increases in the global soil
respiration record. Nature 464: 579-582. http://dx.doi.org/10.1038/nature08930.
Bondi. CA; Beier. CM; Ducev. PK; Lawrence. GB; Bailey. S. (2016). Can the eastern red-
backed salamander (Plethodon cinereus) persist in an acidified landscape? Ecosphere 7:
e01318 (article number), http://dx.doi.org/10.1002/ecs2.1318.
Bonten, LTC; Groenenberg, JE; Meesenburg, H; de Vries, W. (2011). Using advanced surface
complexation models for modelling soil chemistry under forests: Soiling forest,
Germany. Environ Pollut 159: 2831-2839.
http://dx.doi.Org/10.1016/i.envpol.2011.05.002.
June 2018
R-23
DRAFT: Do Not Cite or Quote

-------
Bonten, LTC; Reinds, GJ; Groenenberg, JE; de Vries, W; Posch, M; Evans. CD; Belvazid, S;
Braun. S; Moldan. F; Sverdrup. HU; Kurz. D. (2015). Dynamic geochemical models to
assess deposition impacts and target loads of acidity for soils and surface waters. In W de
Vries; JP Hettelingh; M Posch (Eds.), Critical loads and dynamic risk assessments:
Nitrogen, acidity and metals in terrestrial and aquatic ecosystems (pp. 225-251).
Bilthoven, The Netherlands: Springer Netherlands. http://dx.doi.org/10.1007/978-94-Q17-
9508-1 8.
Boonpragob. K; Nash. T. III. (1991). Physiological responses of the lichen Ramalina Menziesii
Tayl. to the Los Angeles urban environment. Environ Exp Bot 31: 229-238.
http ://dx. doi .org/10.1016/0098-8472(91 N)90075-Y.
Boonpragob, K; Nash, T, III; Fox, CA. (1989). Seasonal deposition patterns of acidic ions and
ammonium to the lichen Ramalina Menziesii Tayl. in Southern California. Environ Exp
Bot 29: 187-197. http://dx.doi.org/10.1016/0098-8472(89^)90051-8.
Boot, CM; Hall, E, dK; Denef, K; Baron, JS. (2016). Long-term reactive nitrogen loading alters
soil carbon and microbial community properties in a subalpine forest ecosystem. Soil
Biol Biochem 92: 211-220. http://dx.doi.Org/10.1016/i.soilbio.2015.10.002.
Booth, MS; Stark, JM; Rastetter, E. (2005). Controls on nitrogen cycling in terrestrial
ecosystems: A synthetic analysis of literature data. Ecol Monogr 75: 139-157.
http://dx.doi.org/10.1890/04-0988.
Borer, ET; Seabloom, EW; Mitchell, CE; Cronin, JP. (2014). Multiple nutrients and herbivores
interact to govern diversity, productivity, composition, and infection in a successional
grassland. Oikos 123: 214-224. http://dx.doi.Org/10.l 111/i. 1600-0706.2013.00680.X.
Borghetti, M; Gentilesca, T; Leonardi, S; vanNoiie, T; Rita, A; Mencuccini, M. (2017). Long-
term temporal relationships between environmental conditions and xylem functional
traits: a meta-analysis across a range of woody species along climatic and nitrogen
deposition gradients. Tree Physiol 37: 4-17. http://dx.doi.org/10.1093/treephys/tpw087.
Boria, A; Basset, A; Bricker, S; Dauvin, JC; Elliott, M; Harrison, T; Marques, JC; Weisberg, S;
West, R. (2012). 1.08: Classifying ecological quality and integrity of estuaries. In D
McLusky (Ed.), Treatise on Estuarine and Coastal Science (pp. 125-162). Amsterdam,
The Netherlands: Academic Press. http://dx.doi.org/10.1016/B978-0-12-374711-2.00109-
L
Boria, A; Dauer, DM; Diaz, R; Llanso, RJ; Muxika, I; Rodriguez, JG; Schaffner, L. (2008).
Assessing estuarine benthic quality conditions in Chesapeake Bay: A comparison of three
indices. Ecol Indicat 8: 395-403. http://dx.doi.Org/10.1016/i.ecolind.2007.05.003.
Borken, W; Beese, F. (2002). Long-term reduction in nitrogen and proton inputs did not affect
atmospheric methane uptake and nitrous oxide emission from a German spruce forest
soil. Soil Biol Biochem 34: 1815-1819.
Borken, W; Matzner, E. (2009). Reappraisal of drying and wetting effects on C and N
mineralization and fluxes in soils. Global Change Biol 15: 808-824.
http://dx.doi.Org/10.llll/i.1365-2486.2008.01681.x.
June 2018
R-24
DRAFT: Do Not Cite or Quote

-------
Bormann. FH; Likens. GE; Fisher. DW; Pierce. RS. (1968). Nutrient loss accelerated by clear-
cutting of a forest ecosystem. Science 159: 882-884.
http://dx.doi.ore/10.1126/science. 159.3817.882.
Bormann. FH: Likens. GE: Melillo. JM. (1977). Nitrogen budget for an aggrading northern
hardwood forest ecosystem. Science 196: 981-983.
http://dx.doi.Org/10.l 126/science. 196.4293.981.
Bosela. M; Petras. R; Sitkova. Z; Priwitzer. T; Paitik. J: Hlavata. Ft; Sedmak. R; Tobin. B.
(2014). Possible causes of the recent rapid increase in the radial increment of silver fir in
the Western Carpathians. Environ Pollut 184: 211-221.
http://dx.doi.Org/10.1016/i.envpol.2013.08.036.
Boshoff. M; De Jonge. M; Dardenne. F; Blust. R; Bervoets. L. (2014). The impact of metal
pollution on soil faunal and microbial activity in two grassland ecosystems. Environ Res
134: 169-180. http://dx.doi.Org/10.1016/i.envres.2014.06.024.
Bott. T; Meyer. GA; Young. EB. (2008). Nutrient limitation and morphological plasticity of the
carnivorous pitcher plant Sarracenia purpurea in contrasting wetland environments. New
Phytol 180: 631-641. http://dx.doi.org/10.1111/i. 1469-8137.2008.02575.x.
Boulton, AJ. (2007). Hyporheic rehabilitation in rivers: Restoring vertical connectivity. Freshw
Biol 52: 632-650. http://dx.doi.Org/10.l 111/i. 1365-2427.2006.01710.X.
Bouskill. NJ: Eveillard. D; Chien. D; Javakumar. A: Ward. BB. (2012). Environmental factors
determining ammonia-oxidizing organism distribution and diversity in marine
environments. Environ Microbiol 14: 714-729. http://dx.doi.Org/10.l 11 l/j.1462-
2920.2011.02623.x.
Bouskill. NJ: Riley. WJ: Tang. JY. (2014). Meta-analysis of high-latitude nitrogen-addition and
warming studies implies ecological mechanisms overlooked by land models.
Biogeosciences 11: 6969-6983. http://dx.doi.org/10.5194/bg-l 1-6969-2014.
Bouwman. AF; Beusen. AHW: Griffioen. J: Van Groenigen. JW: Hefting. MM: Oenema. O:
Van Puiienbroek. PJT. M; Seitzinger. S: Slomp. CP: Stehfest. E. (2013). Global trends
and uncertainties in terrestrial denitrification and N20 emissions. Philos Trans R Soc
Lond B Biol Sci 368: 20130112. http://dx.doi.org/10.1098/rstb.2013.0112.
Bouwman. AF: Lee. DS: Asman. WAH; Dentener. FJ; Vanderhoek. KW: Olivier. JGJ. (1997). A
global high-resolution emission inventory for ammonia. Global Biogeochem Cycles 11:
561-587. http://dx.doi.org/10.1029/97GB02266.
Bowen. JL; Byrnes. JE; Weisman. D; Colaneri. C. (2013). Functional gene pyrosequencing and
network analysis: an approach to examine the response of denitrifying bacteria to
increased nitrogen supply in salt marsh sediments. FMICB 4: 342.
http://dx.doi.org/10.3389/fmicb.2013.00342.
Bowen. JL: Ramstack. J. oyM; Mazzilli. S: Valielai. I. (2007). NLOAD: An interactive, web-
based modeling tool for nitrogen management in estuaries. Ecol Appl 17: S17-S30.
http://dx.doi.Org/10.1890/05-1460.l.
June 2018
R-25
DRAFT: Do Not Cite or Quote

-------
Bo wen. JL; Valiela. I. (2001). The ecological effects of urbanization of coastal watersheds:
Historical increases in nitrogen loads and eutrophication of Waquoit Bay estuaries. Can J
Fish Aquat Sci 58: 1489-1500. http://dx.doi.ore/10.1139/cjfas-58-8-1489.
Bowman. WD. (1994). Accumulation and use of nitrogen and phosphorus following fertilization
in two alpine tundra communities. Oikos 70: 261-270. http://dx.doi.org/10.2307/3545637.
Bowman. WD: Baron. JS: Geiser. LH; Fenn. ME: Lilleskov. EA. (2011). Northwestern forested
mountains. In Assessment of Nitrogen Deposition Effects and Empirical Critical Loads of
Nitrogen for Ecoregions of the United States (pp. 75-88). (NRS-80). Newtown Square,
PA: U.S. Department of Agriculture, Forest Service, Northern Research Station.
http://www.nrs.fs.fed.us/pubs/38109.
Bowman. WD: Fisk. MC. (2001). Chapter 9: Primary production. In WD Bowman; TR Seastedt
(Eds.), Structure and Function of an Alpine Ecosystem: Niwot Ridge, Colorade (pp. 177-
197). Oxford, UK: Oxford University Press.
Bowman. WD: Gartner. JR; Holland. K; Wiedemann. MM. (2006). Nitrogen critical loads for
alpine vegetation and terrestrial ecosystem response: Are we there yet? Ecol Appl 16:
1183-1193. http://dx.doi.org/10.1890/1051-0761(2006^)016T1183:NCLFAV12.0.CQ;2.
Bowman, WD: Murgel, J: Blett, T; Porter, E. (2012). Nitrogen critical loads for alpine vegetation
and soils in Rocky Mountain National Park. J Environ Manage 103: 165-171.
http://dx.doi.Org/10.1016/i.ienvman.2012.03.002.
Bowman, WD: Nemergut, PR: Mcknight, DM: Miller, MP: Williams, MW. (2014). A slide
down a slippery slope - Alpine ecosystem responses to nitrogen deposition. Plant Ecol
Divers 8: 727-738. http://dx.doi.org/10.1080/17550874.2Q14.984786.
Bowman, WD: Steltzer, H. (1998). Positive feedbacks to anthropogenic nitrogen deposition in
Rocky Mountain alpine tundra. Ambio 27: 514-517.
Bowman, WD: Theodose, TA: Fisk, MC. (1995). Physiological and production responses of
plant growth forms to increases in limiting resources in alpine tundra: Implications for
differential community response to environmental change. Oecologia 101: 217-227.
http://dx.doi.org/10.1007/BF0Q317287.
Bowman, WD: Theodose, TA: Schardt, JC: Conant, RT. (1993). Constraints of nutrient
availability on primary production in two alpine tundra communities. Ecology 74: 2085-
2097. http://dx.doi.org/10.2307/1940854.
Boxman, AW: Blanck, K: Brandrud, TE: Emmett, BA: Gundersen, P: Hogervorst, RF: Ki0nass,
OJ: Persson, H: Timmermann, V. (1998a). Vegetation and soil biota response to
experimentally-changed nitrogen inputs in coniferous forest ecosystems of the NITREX
project. For Ecol Manage 101: 65-79. http://dx.doi.org/10.1016/S0378-1127(97N)00126-6.
Boxman, AW: Van der Yen, PJM: Roelofs, JGM. (1998b). Ecosystem recovery after a decrease
in nitrogen input to a Scots pine stand at Ysselsteyn, the Netherlands. For Ecol Manage
101: 155-163. http://dx.doi.org/10.1016/S0378-l 127(97^)00132-1.
June 2018
R-26
DRAFT: Do Not Cite or Quote

-------
Bovce. RL; Schaberg, PG; Hawlev, GJ; Halman. JM; Murakami. PF. (2013). Effects of soil
calcium and aluminum on the physiology of balsam fir and red spruce saplings in
northern New England. Trees Struct Funct 27: 1657-1667.
http ://dx. doi .org/10.1007/s00468-013 -0913 -x.
Boyd. J; Banzhaf. S. (2007). What are ecosystem services? The need for standardized
environmental accounting units. Ecol Econ 63: 616-626.
Boyer. EW; Goodale. CL; Jaworski. NA; Howarth. RW. (2002). Anthropogenic nitrogen sources
and relationships to riverine nitrogen export in the northeastern USA. Biogeochemistry
57: 137-169. http://dx.doi.Org/10.1023/A:1015709302073.
Boyer. KE; Zedler. JB. (1999). Nitrogen addition could shift plant community composition in a
restored California salt marsh. Restor Ecol 7: 74-85. http://dx.doi.org/10.1046/i. 1526-
100X. 1999.07109.x.
Boynton, WR; Hagy, JD; Cornwell, JC; Kemp, WM; Greene, SM; Owens; Baker, JE; Larsen,
RK. (2008). Nutrient budgets and management actions in the Patuxent River Estuary,
Maryland. Estuaries Coasts 31: 623-651. http://dx.doi.org/10.1007/sl2237-008-9Q52-9.
Boynton, WR: Hodgkins, CLS: O'Leary, CA; Bailey, EM: Bayard, AR; Wainger, LA. (2014).
Multi-decade responses of a tidal creek system to nutrient load reductions: Mattawoman
Creek, Maryland USA. Estuaries Coasts 37: S111-S127.
http://dx.doi.org/10.1007/sl2237-013-969Q-4.
Boynton, WR: Kemp, WM. (2008). Chapter 18: Estuaries. In DG Capone; DA Bronk; MR
Mulholland; EJ Carpenter (Eds.), Nitrogen in the Marine Environment (2nd ed., pp. 809-
866). Burlington, MA: Academic Press. http://dx.doi.org/10.1016/B978-Q-12-372522-
6.00018-9.
Bradford, DF; Gordon, MS. (1992). Aquatic amphibians in the Sierra Nevada: Current status and
potential effects of acidic deposition on populations. (ARB/R-93/480). Sacramento, CA:
California University, Los Angelos; Calfornia State Air Resources Board, Research
Division.
https://ntrl.ntis.gov/NTRL/dashboard/searchResults/titleDetail/PB94104288.xhtml.
Bradford, MA; Wood, SA; Maestre, FT; Reynolds, JF; Warren, RJ, II. (2012). Contingency in
ecosystem but not plant community response to multiple global change factors. New
Phytol 196: 462-471. http://dx.doi.Org/10.llll/i.1469-8137.2012.04271.x.
Bradley, K; Driiber, RA; Knops, J. (2006). Increased N availability in grassland soils modifies
their microbial communities and decreases the abundance of arbuscular mycorrhizal
fungi. Soil Biol Biochem 38: 1583-1595. http://dx.doi.Org/10.1016/i.soilbio.2005.l 1.011.
Brady, NC: Weil, RR. (1999). The Nature and Properties of Soils. In The Nature and Properties
of Soils (12th ed.). Upper Saddle River, NJ: Prentice Hall.
Brahnev, J: Mahowald, N: Ward, DS: Ballantyne, AP; Neff, JC. (2015). Is atmospheric
phosphorus pollution altering global alpine Lake stoichiometry? Global Biogeochem
Cycles 29: 1369-1383. http://dx.doi.org/10.1002/2015GB0Q5137.
June 2018
R-27
DRAFT: Do Not Cite or Quote

-------
Branfireun. BA; Roulet NT; Kelly. CA; Rudcl. JWM. (1999). In situ sulphate stimulation of
mercury methylation in a boreal peatland: Toward a link between acid rain and
methylmercury contamination in remote environments. Global Biogeochem Cycles 13:
743-750. http://dx.doi.org/10.1029/1999GB900Q33.
Brauer. VS: Stomp. M; Huisman. J. (2012). The nutrient-load hypothesis: Patterns of resource
limitation and community structure driven by competition for nutrients and light. Am Nat
179: 721-740. http://dx.doi.org/10.1086/665650.
Braun. EL. (2001). Deciduous Forests of Eastern North America. In Deciduous Forests of
Eastern North America. Caldwell, NJ: The Blackburn Press.
Braun. S: Schindler. C: Volz. R; Fltickiger. W. (2003). Forest damages by the storm Lothar in
permanent observation plots in Switzerland: The significance of soil acidification and
nitrogen deposition. Water Air Soil Pollut 142: 327-340.
http://dx.doi.Org/10.1023/A:1022088806060.
Braun. S: Thomas. VF; Quiring. R; Fltickiger. W. (2010). Does nitrogen deposition increase
forest production? The role of phosphorus. Environ Pollut 158: 2043-2052.
http://dx.doi.Org/10.1016/i.envpol.2009.l 1.030.
Breiner, J; Gimeno, BS; Fenn, M. (2007). Calculation of theoretical and empirical nutrient N
critical loads in the mixed conifer ecosystems of southern California.
ScientificWorldJournal 7: 198-205. http://dx.doi.org/10.1100/tsw.20Q7.65.
Breitburg. PL: Hondorp. D; Audemard. C: Carnegie. RB; Burrell. RB; Trice. M; Clark. V.
(2015). Landscape-level variation in disease susceptibility related to shallow-water
hypoxia. PLoS ONE 10: eO 116223. http://dx.doi.org/10.1371/iournal.pone.0116223.
Breitburg. PL: Hondorp. DW: Davias. LA: Diaz. RJ. (2009). Hypoxia, nitrogen, and fisheries:
Integrating effects across local and global landscapes [Review], Ann Rev Mar Sci 1: 329-
349. http://dx.doi.Org/10.l 146/annurev.marine.010908.163754.
Breuninger. C: Oswald. R; Kesselmeier. J: Meixner. FX. (2012). The dynamic chamber method:
Trace gas exchange fluxes (NO, N02, 0-3) between plants and the atmosphere in the
laboratory and in the field. Atmos Meas Tech 5: 955-989. http://dx.doi.org/10.5194/amt-
5-955-2012.
Bricker. OP: Rice. KC. (1989). Acidic deposition to streams: A geology-based method predicts
their sensitivity. Environ Sci Technol 23: 379-385.
http://dx.doi.org/10.1021/esQ0181a001.
Bricker. S: Longstaff. B; Pennison. W: Jones. A: Boicourt. K; Wicks. C: Woerner. J. (2007).
Effects of nutrient enrichment in the nation's estuaries: A decade of change: National
estuarine eutrophication update. Silver Spring, MP: National Centers for Coastal Ocean
Science.
https://vosemite.epa.gov/oa/EAB Web Pocket.nsf/(Filings)/3BE82A42C7EP8C358525
7B120059CB8A/$File/Opposition%20to%20Petition%20for%20Review%20—
%20Ex.%2010%20Partl.. .23.53 .pdf.
Bricker. SB: Clement. CG: Pirhalla. PE; Orlando. SP; Farrow. PGG. (1999). National estuarine
eutrophi cation assessment: Effects of nutrient enrichment in the nation's estuaries. "Silver
Spring, MP: National Oceanic and Atmospheric Administration.
June 2018
R-28
DRAFT: Do Not Cite or Quote

-------
Bricker. SB; Ferreira. JG; Simas. T. (2003). An integrated methodology for assessment of
estuarine trophic status. Ecol Modell 169: 39-60. http://dx.doi.org/ 10.1016/S03 04-
3800(03)00199-6.
Bricker. SB: Longstaf. B; Dennison. W: Jones. A: Boicourt. K; Wicks. C: Woerner. J. (2008).
Effects of nutrient enrichment in the nation's estuaries: A decade of change. Harmful
Algae 8: 21-32. http://dx.doi.Org/10.1016/i.hal.2008.08.028.
Bricker. SB: Rice. KC: Bricker. OP. III. (2014). From headwaters to coast: Influence of human
activities on water quality of the Potomac River estuary. Aquatic Geochemistry 20: 291-
323. http://dx.doi.org/10.1007/slQ498-014-9226-v.
Brigham. ME: Sandheinrich. MB: Gay. DA: Maki. RP; Krabbenhoft. DP: Wiener. JG. (2014).
Lacustrine responses to decreasing wet mercury deposition rates-results from a case study
in northern Minnesota. Environ Sci Technol 48: 6115-6123.
http://dx.doi.org/10.1021/es50Q301a.
Brimblecombe. P; Clegg. SL. (1988). The solubility and behavior of acid gases in the marine
aerosol. J Atmos Chem 7: 1-18. http://dx.doi.org/ 10.1007/BF00048251.
Brin. L; Valiela. I; Goehringer. D; Howes. B. (2010). Nitrogen interception and export by
experimental salt marsh plots exposed to chronic nutrient addition. Mar Ecol Prog Ser
400: 3-17. http://dx.doi.org/10.3354/meps08460.
Bristow. LA; Sarode. N; Cartee. J; Caro-Quintero. A; Thamdrup. B; Stewart. FJ. (2015).
Biogeochemical and metagenomic analysis of nitrite accumulation in the Gulf of Mexico
hypoxic zone. Limnol Oceanogr 60: 1733-1750. http://dx.doi.org/10.1002/lno.1013Q.
Britton. AJ; Fisher. JM. (2007). Interactive effects of nitrogen deposition, fire and grazing on
diversity and composition of low-alpine prostrate Calluna vulgaris heathland. J Appl Ecol
44: 125-135. http://dx.doi.Org/10.llll/i.1365-2664.2006.01251.x.
Britton. AJ; Fisher. JM. (2008). Growth responses of low-alpine dwarf-shrub heath species to
nitrogen deposition and management. Environ Pollut 153: 564-573.
http://dx.doi.Org/10.1016/i.envpol.2007.09.022.
Britton. AJ; Fisher. JM. (2010). Terricolous alpine lichens are sensitive to both load and
concentration of applied nitrogen and have potential as bioindicators of nitrogen
deposition. Environ Pollut 158: 1296-1302.
http://dx.doi.Org/10.1016/i.envpol.2010.01.015.
Britton. AJ; Helliwell. RC; Fisher. JM; Gibbs. S. (2008). Interactive effects of nitrogen
deposition and fire on plant and soil chemistry in an alpine heathland. Environ Pollut
156: 409-416. http://dx.doi.org/10.1016/i.envpol.2008.01.029.
Britton, KO; Berrang, P; Mavity, E. (1996). Effects of pretreatment with simulated acid rain on
the severity of dogwood anthracnose. Plant Dis 80: 646-649.
http://dx.doi.org/10.1094/PD-80-0646.
Brochier-Armanet. C; Boussau. B; Gribaldo. S; Forterre. P. (2008). Mesophilic crenarchaeota:
Proposal for a third archaeal phylum, the Thaumarchaeota. Nat Rev Microbiol 6: 245-
252. http://dx.doi.org/10.1038/nrmicrol852.
June 2018
R-29
DRAFT: Do Not Cite or Quote

-------
Broderick. SJ; Cullen. P; Maher. W. (1988). Denitrification in a natural wetland receiving
secondary treated effluent. Water Res 22: 431-439. http://dx.doi.org/10.1016/0Q43-
1354(88)90037-1.
Brodo. IM: Sharnoff DS: Sharnoff S. (2001). Lichens of North America. New Haven, CT: Yale
University Press.
Bronk. DA: Loin as. MW; Glibert. PM: Schukert. KJ: Sanderson. MP. (2000). Total dissolved
nitrogen analysis: Comparisons between the persulfate, UV and high temperature
oxidation methods. 69: 163-178. http://dx.doi.org/10.1016/S0304-4203(99)00103-6.
Brooks, ML. (1999). Habitat invasibility and dominance by alien annual plants in the western
Mojave desert. Biological Invasions 1: 325-337.
http://dx.doi.Org/10.1023/A:1010057726056.
Brooks. ML. (2003). Effects of increased soil nitrogen on the dominance of alien annual plants in
the Mojave Desert. J Appl Ecol 40: 344-353. http://dx.doi.org/10.1046/j. 1365-
2664.2003.00789.x.
Brooks. ML: D'Antonio. CM: Richardson. DM: Grace. JB; Keelev. JE; DiTomaso. JM; Hobbs.
RJ: Pellant. M; Pyke. D. (2004). Effects of invasive alien plants on fire regimes.
Bioscience 54: 677-688.
Brooks. ML: Esque. TC. (2002). Alien annual plants and wildfire in desert tortoise habitat:
Status, ecological effects, and management. In Alien plants and fire in desert tortoise
(Gopherus agassizii) habitat of the Mojave and Colorado Deserts. Washington, DC: U.S.
Department of the Interior | U.S. Geological Survey (USGS).
https://pubs.er.usgs.gov/publication/1008328.
Brost. RA: Delanv. AC: Huebert. BJ. (1988). NUMERICAL MODELING OF
CONCENTRATIONS AND FLUXES OF HN03, NH3, AND NH4N03 NEAR THE
SURFACE. J Geophys Res Atmos 93: 7137-7152.
Brown. CA: Ozretich. RJ. (2009). Coupling Between the Coastal Ocean and Yaquina Bay,
Oregon: Importance of Oceanic Inputs Relative to Other Nitrogen Sources. Estuaries
Coasts 32: 219-237. http://dx.doi.org/10.1007/sl2237-0Q8-9128-6.
Brown. CA: Power. JH. (2011). Historic and recent patterns of dissolved oxygen in the Yaquina
Estuary (Oregon, USA): Importance of anthropogenic activities and oceanic conditions.
Estuar Coast Shelf Sci 92: 446-455. http://dx.doi.Org/10.1016/i.ecss.2011.01.018.
Brown. DJA. (1983). Effect of calcium and aluminum concentrations on the survival of brown
trout (Salmo trutta) at low pH. Bull Environ Contam Toxicol 30: 582-587.
http://dx.doi.org/10.1007/BF0161Q178.
Brown, RE; Nelson, SJ: Saros, JE. (2017). Paleolimnological evidence of the consequences of
recent increased dissolved organic carbon (DOC) in lakes of the northeastern USA. J
Paleolimnol 57: 19-35. http://dx.doi.org/10.1007/slQ933-016-9913-3.
Bruesewitz, DA: Gardner, WS: Moonev, R, aeF: Pollard, L: Buskev, EJ. (2013). Estuarine
ecosystem function response to flood and drought in a shallow, semiarid estuary:
Nitrogen cycling and ecosystem metabolism. Limnol Oceanogr 58: 2293-2309.
June 2018
R-30
DRAFT: Do Not Cite or Quote

-------
Brunei J: Diekmann. M; Falkengren-Grerup, U. (1998). Effects of nitrogen deposition on field
layer vegetation in south Swedish oak forests. Environ Pollut 102: 35-40.
http ://dx. doi .org/10.1016/S0269-7491 (98)80012-2.
Brush. GS. (2009). Historical land use, nitrogen, and coastal eutrophication: A paleoecological
perspective. Estuaries Coasts 32: 18-28. http://dx.doi.org/10.1007/sl2237-008-91Q6-z.
Bubier. JL; Smith. R; Juutinen. S: Moore. T. imR; Minocha. R; Long. S: Minocha. S. (2011).
Effects of nutrient addition on leaf chemistry, morphology, and photosynthetic capacity
of three bog shrubs. Oecologia 167: 355-368. http://dx.doi.org/10.1007/sQ0442-011-
1998-9.
Buckler. PR: Mehrle. PM; Cleveland. L; Dwver. FJ. (1987). Influence of pH on the toxicity of
aluminum and other inorganic contaminants to East Coast striped bass. Water Air Soil
Pollut 35: 97-106. http://dx.doi.org/10.1007/BF0Q183846.
Bucsela, EJ; Krotkov, NA; Celarier, EA; Lamsal, LN; Swartz, WH; Bhartia, PK; Boersma, KF;
Veefkind. JP; Gleason. JF; Pickering. KE. (2013). A new stratospheric and tropospheric
N02 retrieval algorithm for nadir-viewing satellite instruments: Applications to OMI.
Atmospheric Measurement Techniques Discussions 6: 2607-2626.
http://dx.doi.org/10.5194/amtd-6-1361-2013.
Buda. AR; DeWalle. DR. (2009). Dynamics of stream nitrate sources and flow pathways during
stormflows on urban, forest and agricultural watersheds in central Pennsylvania, USA.
Hydrolog Process 23: 3292-3305. http://dx.doi.org/10.1002/hyp.7423.
Buechling, A; Baker, WL. (2004). A fire history from tree rings in a high-elevation forest of
Rocky Mountain National Park. Can J For Res 34: 1259-1273.
http ://dx.doi.org/10.113 9/X04-012.
Bukaveckas. PA: Lesutiene. J: Gasiunaite. ZR; Lozys. L; Olenina. I: Pilkaitvte. R; Putvs. Z;
Tassone, S; Wood, J. (2017). Microcystin in aquatic food webs of the Baltic and
Chesapeake Bay regions. Estuar Coast Shelf Sci 191: 50-59.
http://dx.doi.Org/10.1016/i.ecss.2017.04.016.
Bulger, AJ: Cosby, BJ; Dolloff, CA; Eshleman, KN: Webb, JR; Galloway, JN. (1999).
SNP:FISH. Shenandoah National Park: Fish in sensitive habitats. Project final report-
Volume 1-4 (pp. 1-152). Charlottesville, VA: University of Virginia.
Bulger, AJ: Cosby, BJ: Webb, JR. (2000). Current, reconstructed past, and projected future status
of brook trout (Salvelinus fontinalis) streams in Virginia. Can J Fish Aquat Sci 57: 1515-
1523. http://dx.doi.Org/10.l 139/cjfas-57-7-l515.
Burd, AB; Frev, S: Cabre, A: Ito, T; Levine, NM; L0nborg, C: Long, M; Mauritz, M; Thomas,
RQ: Stephens, BM; Vanwalleghem, T; Zeng, N. (2016). Terrestrial and marine
perspectives on modeling organic matter degradation pathways [Review], Global Change
Biol 22: 121-136. http://dx.doi.org/10.1111/gcb. 12987.
Burgmer, T; Hillebrand, Ft; Pfenninger, M. (2007). Effects of climate-driven temperature
changes on the diversity of freshwater macroinvertebrates. Oecologia 151: 93-103.
http://dx.doi.org/10.1007/s00442-006-Q542-9.
June 2018
R-31
DRAFT: Do Not Cite or Quote

-------
Burkhardt J: Basi. S; Pariyar. S; Hunsche. M. (2012). Stomatal penetration by aqueous solutions
- an update involving leaf surface particles. New Phytol 196: 774-787.
http://dx.doi.org/10.111 l/i.l469-8137.2012.04307.x.
Burkhardt J: Flechard. CR; Gresens. F; Mattsson. ME: Jongeian. PAC: Erisman. JW:
Wei dinger. T; Meszaros. R; Nemitz. E: Sutton. MA. (2008). Modeling the dynamic
chemical interactions of atmospheric ammonia and other trace gases with measured leaf
surface wetness in a managed grassland canopy. Biogeosciences Discussions 5: 2505-
2540.
Burkhardt. J: Pariyar. S. (2014). Particulate pollutants are capable to 'degrade' epicuticular waxes
and to decrease the drought tolerance of Scots pine (Pinus sylvestris L.). Environ Pollut
184: 659-667. http://dx.doi.org/10.1016/i.envpol.2013.04.041.
Burkhardt. J: Pariyar. S. (2015). How does the VPD response of isohydric and anisohydric plants
depend on leaf surface particles? Plant Biol (Stuttg) 18: 91-100.
http://dx.doi.org/10. Ill 1/plb. 12402.
Burnell. OW: Russell. BP: Irving. AD: Connell. SD. (2013). Eutrophication offsets increased sea
urchin grazing on seagrass caused by ocean warming and acidification. Mar Ecol Prog
Ser485: 37-46. http://dx.doi.org/10.3354/mepsl0323.
Burns. DA. (1996). The effects of liming an Adirondack lake watershed on downstream water
chemistry. Biogeochemistry 32: 339-362. http://dx.doi.org/10.1007/BFQ2187145.
Burns. DA: Blett. T; Haeuber. R; Pardo. LH. (2008a). Critical loads as a policy tool for
protecting ecosystems from the effects of air pollutants. Front Ecol Environ 6: 156-159.
http://dx.doi.org/10.1890/07004Q.
Burns. DA: Bover. EW: Elliott. EM: Kendall. C. (2009). Sources and transformations of nitrate
from streams draining varying land uses: Evidence from dual isotope analysis. J Environ
Qual 38: 1149-1159. http://dx.doi.org/10.2134/ieq2008.0371.
Burns. DA: Riva-Murrav. K; Bode. RW: Passv. S. (2008b). Changes in stream chemistry and
biology in response to reduced levels of acid deposition during 1987-2003 in the
Neversink River Basin, Catskill Mountains. Ecol Indicat 8: 191-203.
http://dx.doi.Org/10.1016/i.ecolind.2007.01.003.
Burrows. RM: Hotchkiss. ER: Jonsson. M: Laudon. Ft: McKie. BG: Sponseller. RA. (2015).
Nitrogen limitation of heterotrophic biofilms in boreal streams. Freshw Biol 60: 1237-
1251. http://dx.doi.org/10. Ill 1/fwb. 12549.
Burton. AJ: Jarvev. JC: Jarvi. MP: Zak. PR: Pregitzer. KS. (2012). Chronic N deposition alters
root respiration-tissue N relationship in northern hardwood forests. Global Change Biol
18: 258-266. http://dx.doi.org/10.1111/i. 1365-2486.2011,02527.x.
Businger. JA. (1986). Evaluation of the accuracy with which dry deposition can be measured
with current micrometeorological techniques. J Clim Appl Meteorol 25: 1100-1124.
Butler. T; Vermevlen. F; Lehmann. CM: Likens. GE: Puchalski. M. (2016). Increasing ammonia
concentration trends in large regions of the USA derived from the NADP/AMoN
network. Atmos Environ 146: 132-140.
http://dx.doi.Org/10.1016/i.atmosenv.2016.06.033.
June 2018
R-32
DRAFT: Do Not Cite or Quote

-------
Butterbach-Bahl. K; Nemitz. E; Zaehle. S; Bill en, G; Boeckx. P; Erisman. JW; Gamier. J;
Upstill-Goddard, R; Kreuzer, M; Oenema, O; Reis, S; Schaap, M; Simpson. D; de Vries,
W; Winiwarter, W; Sutton, MA. (2011). Nitrogen as a threat to the European greenhouse
balance. In The European Nitrogen Assessment: Sources, Effects and Policy
Perspectives. Cambridge, UK: Cambridge University Press.
Buzorius, G: Rannik, U; Makela, JM; Vesala, T; Kulmala, M. (1998). Vertical aerosol particle
fluxes measured by eddy covariance technique using condensational particle counter. J
Aerosol Sci 29: 157-171. http://dx.doi.org/10.1016/S0021-8502(97N)00458-8.
Buzzelli, C: Chen, Z; Colev, T; Doering, P; Samimv, R; Schlezinger, D; Howes, B. (2013a). Dry
season sediment-water exchanges of nutrients and oxygen in two Florida estuaries:
Patterns, comparisons, and internal loading. Florida Scientist 76: 54-79.
Buzzelli, C: Doering, PH: Wan, Y; Sun, D; Fugate, D. (2014). Modeling ecosystem processes
with variable freshwater inflow to the Caloosahatchee River Estuary, southwest Florida.
I. Model development. Estuar Coast Shelf Sci 151: 256-271.
http://dx.doi.Org/10.1016/i.ecss.2014.08.028.
Buzzelli, C: Wan, Y; Doering, PH: Bover, JN. (2013b). Seasonal dissolved inorganic nitrogen
and phosphorus budgets for two sub-tropical estuaries in south Florida, USA.
Biogeosciences 10: 6721-6736. http://dx.doi.org/10.5194/bg-10-6721-2013.
Bytnerowicz, A: Fenn, M; Mcnultv, S: Yuan, F; Pourmokhtarian, A: Driscoll, C: Meixner, T,
om. (2013). Interactive effects of air pollution and climate change on forest ecosystems in
the United States: Current understanding and future scenarios. In Developments in
Environmental Science. Amsterdam, Netherlands: Elsevier Ltd.
Bytnerowicz, A: Fenn, ME. (1996). Nitrogen deposition in California forests: A review
[Review], Environ Pollut 92: 127-146. http://dx.doi.org/10.1016/0269-7491(95^)00106-9.
Bytnerowicz, A: Johnson, RF; Zhang, L; Jenerette, GD; Fenn, ME: Schilling, SL; Gonzalez-
Fernandez, I. (2015). An empirical inferential method of estimating nitrogen deposition
to Mediterranean-type ecosystems: The San Bernardino Mountains case study. Environ
Pollut 203: 69-88. http://dx.doi.Org/10.1016/i.envpol.2015.03.028.
Bytnerowicz, A: Percy, K; Riechers, G: Padgett, P; Krywult, M. (1998). Nitric acid vapor effects
on forest trees - Deposition and cuticular changes. Chemosphere 36: 697-702.
http://dx.doi.org/10.1016/S0045-6535(97N)10110-2.
Byun, D; Schere, KL. (2006). Review of the governing equations, computational algorithms, and
other components of the models-3 community multiscale air quality (CMAQ) modeling
system [Review], Appl Mech Rev 59: 51-77. http://dx.doi.Org/10.l 115/1.2128636.
Byun, DW: Ching, JKS. (1999). Science algorithms of the EPA models-3 community multiscale
air quality (CMAQ) modeling system. (EPA/600-R-99-030). Washington, DC: U.S.
Environmental Protection Agency.
http://www.epa.gov/asmdnerl/CMAQ/CMAQscienceDoc.html.
CAA. Clean Air Act, as amended by Pub. L. No. 101-549, section 108: Air quality criteria and
control techniques, 42 USC 7408, 42 USC (1990a).
http://www.law.cornell.edu/uscode/text/42/7408
June 2018
R-33
DRAFT: Do Not Cite or Quote

-------
CAA (Clean Air Act). (1990b). Clean Air Act, as amended by Pub. L. No. 101-549, section 109:
National primary and secondary ambient air quality standards, 42 USC 7409.
https://www.law.cornell.edu/uscode/text/42/7409.
CAA. Clean Air Act, section 302: Definitions. 42 USC 7602. 42 USC (2005).
http://www.gpo.gOv/fdsvs/pkg/USCODE-2005-title42/pdf/USCODE-2005-title42-
chap 8 5 - sub chapIII- sec7602. pdf
Cabral. L; Fonseca Sousa Spares. CR; Giachini. AJ; Siqueira. JO. (2015). Arbuscular
mycorrhizal fungi in phytoremediation of contaminated areas by trace elements:
mechanisms and major benefits of their applications. World J Microbiol Biotechnol 31:
1655-1664. http://dx.doi.org/10.1007/sll274-Q15-1918-v.
Caddy. JF. (1993). Toward a comparative evaluation of human impacts on fishery ecosystems of
enclosed and semi-enclosed seas. Rev Fish Sci 1: 57-95.
http://dx.doi.org/10.1080/106412693Q9388535.
Caffrev. JM; Murrell. MC: Wigand. C: McKinnev. R. (2007). Effect of nutrient loading on
biogeochemical and microbial processes in a New England salt marsh. Biogeochemistry
82: 251-264. http://dx.doi.org/10.1007/slQ533-007-9068-4.
Cai, J; Weiner, J; Wang, R; Luo, W; Zhang, Y; Liu, H; Xu, Z; Li, H; Zhang, Y; Jiang, Y.
(2017a). Effects of nitrogen and water addition on trace element stoichiometry in five
grassland species. J Plant Res 130: 659-668. http://dx.doi.org/10.1007/slQ265-017-0928-
2.
Cai, M; Johnson, AM; Schwartz, JS; Moore, SE; Kulp, MA. (201 la). Response of soil water
chemistry to simulated changes in acid deposition in the Great Smoky Mountains. J
Environ Eng 137: 617-628. http://dx.doi.org/10.^l/fASCE^EE. 1943-7870.0000354.
Cai, M; Johnson, AM: Schwartz, JS: Moore, SE: Kulp, MA. (2012). Soil acid-base chemistry of
a high-elevation forest watershed in the Great Smoky Mountains National Park: Influence
of acidic deposition. Water Air Soil Pollut 223: 289-303.
http://dx.doi.org/10.1007/sll270-011-Q858-x.
Cai, M; Schwartz, JS: Robinson, RB; Moore, SE: Kulp, MA. (2010). Long-term effects of acidic
deposition on water quality in a high-elevation Great Smoky Mountains National Park
watershed: Use of an ion input-output budget. Water Air Soil Pollut 209: 143-156.
http://dx.doi.org/10.1007/sll270-009-Q187-5.
Cai, M; Schwartz, JS: Robinson, RB: Moore, SE: Kulp, MA. (201 lb). Long-term annual and
seasonal patterns of acidic deposition and stream water quality in a Great Smoky
Mountains high-elevation watershed. Water Air Soil Pollut 219: 547-562.
http://dx.doi.org/10.1007/sll270-010-Q727-z.
Cai, W, -J; Hu, X; Huang, W, -J; Murrell, MC; Lehrter, JC: Lohrenz, SE; Chou, W, -C: Zhai, W:
Hollibaugh, JT; Wang, Y; Zhao, P; Guo, X; Gundersen, K; Dai, M; Gong, G, -C.
(2011c). Acidification of subsurface coastal waters enhanced by eutrophication. Nat
Geosci 4: 766-770. http://dx.doi.org/10.1038/ngeol297.
June 2018
R-34
DRAFT: Do Not Cite or Quote

-------
Cat WJ; Huang. WJ; Luther. GW; Pierrot D: Li. M; Testa. J; Xue. M; Joesoef. A; Mann. R;
Brodeur. J; Xu. YY; Chen. B; Hussain. N; Waldbusser. GG; Cornwell. J; WM. K.
(2017b). Redox reactions and weak buffering capacity lead to acidification in the
Chesapeake Bay. Nat Commun 8: Article #369. http://dx.doi.org/10.1038/s41467-Q17-
00417-7.
Caine. N. (1995). Temporal trends in the quality of streamwater in an Alpine environment:
Green Lakes valley, Colorado Front Rand, USA. Geografiska Annaler Series A: Physical
Geography 77A: 207-220. http://dx.doi.org/10.2307/521330.
Caires. AM: Chandra. S: Havford. BL; Wittmann. ME. (2013). Four decades of change:
Dramatic loss of zoobenthos in an oligotrophic lake exhibiting gradual eutrophication.
Freshw Sci 32: 692-705. http://dx.doi.org/10.1899/12-064.1.
Calanni. J: Berg. E; Wood. M; Mangis. D; Bovce. R; Weathers. W: Sievering. H. (1999).
Atmospheric nitrogen deposition at a conifer forest: Response of free amino acids in
Engelmann spruce needles. Environ Pollut 105: 79-89. http://dx.doi.org/10.1016/SQ269-
7491(98)00202-4.
Cale. JA; Garrison-Johnston. MT; Teale. SA; Castello. JD. (2017). Beech bark disease in North
America: Over a century of research revisited. For Ecol Manage 394: 86-103.
http://dx.doi.Org/10.1016/i.foreco.2017.03.031.
Calhoun. A: Cormier. J: Owen. R; OConnell. A: Roman. C: Tiner. R. (1994). The wetlands of
Acadia National Park and vicinity. (Miscellaneous Publication 721). Maine Agricultural
and Forest Experiment Station, The National Park Service, The U.S. Fish and Wildlife
Service.
Callesen. I; Nilsson. L; Schmidt. IK: Vesterdal. L; Ambus. P; Christiansen. JR; Hogberg. P;
Gundersen. P. (2013). The natural abundance of N-15 in litter and soil profiles under six
temperate tree species: N cycling depends on tree species traits and site fertility. Plant
Soil 368: 375-392. http://dx.doi.org/10.1007/slll04-012-1515-x.
Camargo. JA: Alonso. A. (2006). Ecological and toxicological effects of inorganic nitrogen
pollution in aquatic ecosystems: A global assessment [Review], Environ Int 32: 831-849.
http://dx.doi.Org/10.1016/i.envint.2006.05.002.
Camenzind. T; Hempel. S: Homeier. J: Horn. S: Velescu. A: Wilcke. W: Rillig. MC. (2014).
Nitrogen and phosphorus additions impact arbuscular mycorrhizal abundance and
molecular diversity in a tropical montane forest. Global Change Biol 20: 3646-3659.
http://dx.doi.org/10. Ill 1/gcb. 12618.
Campbell. DH; Muths. E; Turk. JT; Corn. PS. (2004a). Sensitivity to acidification of subalpine
ponds and lakes in north-western Colorado. Hydrolog Process 18: 2817-2834.
http://dx.doi.org/10.1002/hyp.1496.
Campbell, JL; Driscoll, CT; Eagar, C; Likens, GE; Siccama, TG; Johnson, CE; Fahey, TJ;
Hamburg, SP; Holmes, RT; Bailey, AS: Buso, DC. (2007). Long-term trends from
ecosystem research at the Hubbard Brook Experimental Forest. (Gen. Tech. Rep. NRS-
17). Newton Square, PA: U.S. Department of Agriculture, Forest Service, Northern
Research Station, http://www.nrs.fs.fed.us/pubs/3119.
June 2018
R-35
DRAFT: Do Not Cite or Quote

-------
Campbell JL; Hornbeck, JW; Mitchell MJ; Adams. MB; Castro. MS; Driscoll, CT; Kahl, JS;
Kochenderfer. JN; Likens. GE; Lynch. JA; Murdoch. PS; Nelson. SJ; Shanlev. JB.
(2004b). Input-output budgets of inorganic nitrogen for 24 forest watersheds in the
northeastern United States: A review. Water Air Soil Pollut 151: 373-396.
http://dx.doi.Org/10.1023/B:WATE.0000009908.94219.04.
Campbell. JL; Ollinger, SV; Flerchinger, GN; Wicklein. H; Havhoe. K; Bailey. AS. (2010). Past
and projected future changes in snowpack and soil frost at the Hubbard Brook
Experimental Forest, New Hampshire, USA. Hydrolog Process 24: 2465-2480.
http://dx.doi.org/10.1002/hyp.7666.
Campbell. JL; Rustad. LE; Bover. EW; Christopher. SF; Driscoll. CT; Fernandez. IJ; Groffman.
PM; Houle. D; Kiekbusch. J; Magill. AH; Mitchell. MJ; Ollinger. SV. (2009).
Consequences of climate change for biogeochemical cycling in forests of northeastern
North America. Can J For Res 39: 264-284. http://dx.doi.Org/10.l 139/x08-104.
Campbell. LM. (2002). The biology of sea turtles contemporary culture, use, and conservation of
sea turtles. In The Biology of Sea Turtles. Boca Raton, FL: CRC Press.
Candolin. U. (2009). Population responses to anthropogenic disturbance: Lessons from three-
spined sticklebacks Gasterosteus aculeatus in eutrophic habitats. J Fish Biol 75: 2108-
2121. http://dx.doi.org/10.111 l/i.l095-8649.2009.02405.x.
Canty. TP; Hembeck. L; Vinciguerra. TP; Anderson. DC; Goldberg. PL; Carpenter. SF; Allen.
DJ; Loughner. CP; Salawitch. RJ; Dickerson. RR. (2015). Ozone and NOx chemistry in
the eastern US: evaluation of CMAQ/CB05 with satellite (OMI) data. Atmos Chem Phys
15: 10965-10982. http://dx.doi.org/10.5194/acp-15-10965-2015.
Cape. JN. (2003). Effects of airborne volatile organic compounds on plants [Review], Environ
Pollut 122: 145-157.
Cape, JN; Cornell, SE; Jickells, TP; Nemitz, E. (2011). Organic nitrogen in the atmosphere
Where does it come from? A review of sources and methods. Atmos Res 102: 30-48.
http://dx.doi.Org/10.1016/i.atmosres.2011.07.009.
Caplan, JS; Wheaton, CN; Mozdzer, TJ. (2014). Belowground advantages in construction cost
facilitate a cryptic plant invasion. 6. http://dx.doi.org/10.1093/aobpla/plu020.
Caputo, J; Beier, CM; Fakhraei, H; Driscoll, CT. (2017). Impacts of Acidification and Potential
Recovery on the Expected Value of Recreational Fisheries in Adirondack Lakes (USA).
Environ Sci Technol 51: 742-750. http://dx.doi.org/10.1021/acs.est.6b05274.
Caraco, NF; Cole, JJ. (1999). Human impact on nitrate export: An analysis using major world
rivers. Ambio 28: 167-170.
Cardinale, BJ; Duffy, JE; Gonzalez, A; Hooper, DU; Perrings, C; Venail, P; Narwani, A; Mace,
GM; Tilman, D; Wardle, DA; Kinzig, AP; Daily, GC; Loreau, M; Grace, JB;
Larigauderie, A; Srivastava, DS; Naeem, S. (2012). Biodiversity loss and its impact on
humanity [Review], Nature 486: 59-67. http://dx.doi.org/10.1038/naturel 1148.
June 2018
R-36
DRAFT: Do Not Cite or Quote

-------
Cardinale. BJ; Matulich. KL; Hooper. DU; Byrnes. JE; Duffy. E; Gamfeldt L; Balvanera. P;
O'Connor. MI; Gonzalez. A. (2011). THE FUNCTIONAL ROLE OF PRODUCER
DIVERSITY IN ECOSYSTEMS. Am J Bot 98: 572-592.
http://dx.doi.org/10.3732/aib.100Q364.
Carey. CJ: Dove. NC: Beman. JM; Hart. SC: Aronson. EL. (2016). Meta-analysis reveals
ammonia-oxidizing bacteria respond more strongly to nitrogen addition than ammonia-
oxidizing archaea. Soil Biol Biochem 99: 158-166.
http://dx.doi.Org/10.1016/i.soilbio.2016.05.014.
Carleton. JN; Park. RA; Clough. JS. (2009). Ecosystem Modeling Applied to Nutrient Criteria
Development in Rivers. Environ Manage 44: 485-492. http://dx.doi.org/10.1007/sQ0267-
009-9344-2.
Carling. GT; Richards. DC: Hoven. H; Miller. T; Fernandez. DP: Rudd. A: Pazmino. E; Johnson.
WP. (2013). Relationships of surface water, pore water, and sediment chemistry in
wetlands adjacent to Great Salt Lake, Utah, and potential impacts on plant community
health. Sci Total Environ 443: 798-811.
http://dx.doi. org/10.1016/i. scitotenv.2012.11.063.
Carlisle. DM: Meador. MR. (2007). A biological assessment of streams in the eastern united
states using a predictive model for macroinvertebrate assemblages. J Am Water Resour
Assoc 43: 1194-1207. http://dx.doi.Org/10.l 111/j. 1752-1688.2007.00097.X.
Carlisle. DM: Meador. MR: Moulton. SR; Ruhl. PM. (2007). Estimation and application of
indicator values for common macroinvertebrate genera and families of the United States.
Ecol Indicat 7: 22-33. http://dx.doi.Org/10.1016/i.ecolind.2005.09.005.
Carlson. RW. (1983). Interaction between S02 and N02 and their effects on photosynthetic
properties of soybean glycine max. Environ Pollut 32: 11-38.
http://dx.doi.org/10.1016/0143-1471(83^)90071-5.
Carmichael. RH; Shriver. AC: Valiela. I. (2004). Changes in shell and soft tissue growth, tissue
composition, and survival of quahogs, Mercenaria mercenaria, and softshell clams, Mya
arenaria, in response to eutrophic-driven changes in food supply and habitat. Exp Mar
Bio Ecol 313: 75-104. http://dx.doi.Org/10.1016/i.iembe.2004.08.006.
Carmichael. RH; Shriver. AC; Valiela. I. (2012). Bivalve response to estuarine eutrophication:
The balance between enhanced food supply and habitat alterations. J Shellfish Res 31:1-
11. http://dx.doi.org/10.2983/035.031.Q101.
Carreiro, MM; Sinsabaugh, RL; Repert, DA; Parkhurst, DF. (2000). Microbial enzyme shifts
explain litter decay responses to simulated nitrogen deposition. Ecology 81: 2359-2365.
http://dx.doi.org/10.1890/0012-9658(2000N)08ir2359:MESELD12.0.CQ:2.
Caruso. NM; Lips. KR. (2013). Truly enigmatic declines in terrestrial salamander populations in
Great Smoky Mountains National Park. Diversity and Distributions 19: 38-48.
http://dx.doi.org/10.111 l/i.l472-4642.2012.00938.x.
Carvalho. LM; Cacador. I; Martins-Loucao. MA. (2006). Arbuscular mycorrhizal fungi enhance
root cadmium and copper accumulation in the roots of the salt marsh plant Aster
tripolium L. Plant Soil 285: 161-169. http://dx.doi.org/10.1007/sl 1104-006-9001-y.
June 2018
R-37
DRAFT: Do Not Cite or Quote

-------
Casson, NJ; Eimers, MC; Watmough, SA. (2012). Impact of winter warming on the timing of
nutrient export from forested catchments. Hydrolog Process 26: 2546-2554.
http://dx.doi.ore/10.1002/hyp.8461.
Casson, NJ: Eimers, MC: Watmough, SA. (2014a). Controls on soil nitrification and stream
nitrate export at two forested catchments. Biogeochemistry 121: 355-368.
http://dx.doi.org/10.1007/sl0533-014-00Q6-v.
Casson, NJ: Eimers, MC: Watmough, SA. (2014b). Sources of nitrate export during rain-on-
snow events at forested catchments. Biogeochemistry 120: 23-36.
http://dx.doi.org/10.1007/slQ533-013-9850-4.
Castro. MS: Driscoll, CT. (2002). Atmospheric nitrogen deposition to estuaries in the mid-
atlantic and northeastern United States. Environ Sci Technol 36: 3242-3249.
http://dx.doi.org/10.1021/es010664o.
Castro, MS: Driscoll, CT; Jordan, TE; Reay, WG: Boynton, WR. (2003). Sources of nitrogen to
estuaries in the United States. Estuaries Coasts 26: 803-814.
http://dx.doi.org/10.1007/BF02711991.
Castro, MS: Driscoll, CT: Jordan, TE: Reay, WG: Boynton, WR: Seitzinger, SP: Styles, RV:
Cable, JE. (2001). Contribution of atmospheric depostition to the total nitrogen loads to
thirty-four estuaries on the Atlantic and Gulf Coasts of the United States. In RA Valigura;
RB Alexander; MS Castro; TP Meyers; HW Paerl; PE Stacey; RE Turner (Eds.),
Nitrogen Loading in Coastal Water Bodies: An Atmospheric Perspective (pp. 77-106).
Washington, DC: American Geophysical Union. http://dx.doi.org/10.1029/CE057pQ077.
Caswell, BA: Frid, CLJ. (2013). Learning from the past: Functional ecology of marine benthos
during eight million years of aperiodic hypoxia, lessons from the late Jurassic. Oikos 122:
1687-1699. http://dx.doi.org/10.111 l/i.l600-0706.2013.00380.x.
Cavlovic, J; Boncina, A; Bozic, M; Gorsic, E; Simoncic, T; Teslak, K. (2015). Depression and
growth recovery of silver fir in uneven-aged Dinaric forests in Croatia from 1901 to
2001. Forestry 88: 586-598. http://dx.doi.org/10.1093/forestry/cpv026.
CDPHE (Colorado Department of Public Health and Environment). (2012). RMNP Initiative
accomplishment timeline. Available online at
https://www.colorado.gov/pacific/sites/default/files/AP PO RMNP-Initiative-
Accomplishment-Timeline.pdf (accessed
Cerco, CF. (2015). A multi-module approach to calculation of oyster (Crassostrea virginica)
environmental benefits. Environ Manage 56: 467-479. http://dx.doi.org/10.1007/sQ0267-
015-0511-3.
Cerco, CF: Noel, MR. (2007). Can oyster restoration reverse cultural eutrophication in
Chesapeake Bay? Estuaries Coasts 30: 331-343. http://dx.doi.org/10.1007/BF0270Q175.
Cerco, CF: Tillman, D: Hagy, JD. (2010). Coupling and comparing a spatially- and temporally-
detailed eutrophication model with an ecosystem network model: An initial application to
Chesapeake Bay. Environ Modell Softw 25: 562-572.
http://dx.doi.Org/10.1016/i.envsoft.2009.09.008.
June 2018
R-38
DRAFT: Do Not Cite or Quote

-------
Cesarz. S; Reich. PB; Scheu. S; Ruess. L; Schaefer. M; Eisenhauer. N. (2015). Nematode
functional guilds, not trophic groups, reflect shifts in soil food webs and processes in
response to interacting global change factors. Pedobiologia 58: 23-32.
http://dx.doi.Org/10.1016/i.pedobi.2015.01.001.
Cha. DH; Appel. HM; Frost CJ: Schultz. JC: Steiner. KC. (2010). Red oak responses to nitrogen
addition depend on herbivory type, tree family, and site. For Ecol Manage 259: 1930-
1937. http://dx.doi.ore/10.1016/i.foreco.2010.02.004.
Chad wick. MA: Huryn. AD. (2005). Response of stream macroinvertebrate production to
atmospheric nitrogen deposition and channel drying. Limnol Oceanogr 50: 228-236.
http://dx.doi.Org/10.4319/lo.2005.50.l.0228.
Chaffin. J: Bridgeman. T; Bade. D. (2013). Nitrogen constrains the growth of late summer
cyanobacterial blooms in Lake Erie. Adv Microbiol 3: 16-26.
http://dx.doi.org/10.4236/aim.2013.36A003.
Chamberlain. AC. (1967). Transport of lycopodium spores and other small particles to rough
surfaces. Proc R Soc Lond A Math Phys Sci 296: 45-70.
Chambers. PL: Woidak. JM; Du. P; Belden. LK. (2013). Pond acidification may explain
differences in corticosterone among salamander populations. Physiol Biochem Zool 86:
224-232. http://dx.doi.org/10.1086/669917.
Chameides. WL; Yu. H: Liu. SC: Bergin. M; Zhou. X: Mearns. L; Wang. G: Kiang. CS: Savior.
RD; Luo. C: Huang. Y; Steiner. A: Giorgi. F. (1999). Case study of the effects of
atmospheric aerosols and regional haze on agriculture: an opportunity to enhance crop
yields in China through emission controls? Proc Natl Acad Sci USA 96: 13626-13633.
Chan. EAW: Gantt. B; Mcdow. S. (2018). The reduction of summer sulfate and switch from
summertime to wintertime PM2.5 concentration maxima in the United States. Atmos
Environ 175: 25-32. http://dx.doi.Org/10.1016/i.atmosenv.2017.l 1.055.
Chan. F; Boehm. AB; Barth. JA; Chornesky. EA; Dickson. AG: Feelv. RA; Hales. B; Hill. TM;
Hofmann. G: Ianson. D; Klinger. T; Largier. J: Newton. J: Pedersen. TF; Somero. GN:
Sutula. M; Wakefield. WW: Waldbusser. GG: Weisberg. SB: Whiteman. EA. (2016).
The West Coast Ocean Acidification and Hypoxia Science Panel: Major findings,
recommendations, and actions. Oakland, CA: California Ocean Science Trust.
http://westcoastoah.org/wp-content/uploads/2016/04/QAH-Panel-Kev-Findings-
Recommendations-and-Actions-4.4.16-FINAL.pdf.
Chandra, S; Vander Zanden, MJ; Heyvaert, AC: Richards, BC; Allen, BC; Goldman, CR.
(2005). The effects of cultural eutrophication on the coupling between pelagic primary
producers and benthic consumers. Limnol Oceanogr 50: 1368-1376.
http://dx.doi.Org/10.4319/lo.2005.50.5.1368.
Chapin, FS, III: Matson, PA; Mooney, HA. (2002). Principles of terrestrial ecosystem ecology.
New York, NY: Springer-Verlag New York, Inc.
Chapin, III, FS. (1980). The mineral nutrition of wild plants. Annu Rev Ecol Evol Systemat 11:
233-260. http://dx.doi.Org/10.l 146/annurev.es.l 1.110180.001313.
June 2018
R-39
DRAFT: Do Not Cite or Quote

-------
Chapman. PJ; Clark. JM; Reynolds. B; Adam son. JK. (2008). The influence of organic acids in
relation to acid deposition in controlling the acidity of soil and stream waters on a
seasonal basis. Environ Pollut 151: 110-120.
http://dx.doi.Org/10.1016/i.envpol.2007.03.001.
Chapman. SK; Devine. KA; Curran. C: Jones. RO: Gilliam. FS. (2016). Impacts of soil nitrogen
and carbon additions on forest understory communities with a long nitrogen deposition
history. Ecosystems 19: 142-154. http://dx.doi.ore/10.1007/sl0021-015-9922-5.
Chappelka. AH: Chevone. BI; Burk. TE. (1988). Growth response of green and white ash
seedlings to ozone, sulfur dioxide, and simulated acid rain. Forest Sci 34: 1016-1029.
Charles. DF; Christie. S. (1991). Acidic deposition and aquatic ecosystems: Regional case
studies. In DF Charles; S Christie (Eds.), Acidic Deposition and Aquatic Ecosystems:
Regional Case Studies. New York, NY: Springer-Verlag. http://dx.doi.org/10.1007/978-
1-4613-9038-1.
Chaturvedi. RK; Prasad. S: Rana. S: Obaidullah. SM; Pandev. V: Singh. H. (2013). Effect of
dust load on the leaf attributes of the tree species growing along the roadside. Environ
Monit Assess 185: 383-391. http://dx.doi.org/10.1007/slQ661-012-2560-x.
Chen, B; Coops, NC; Black, TA; Jassal, RS; Chen, JM; Johnson, M. (2011). Modeling to discern
nitrogen fertilization impacts on carbon sequestration in a Pacific Northwest Douglas-fir
forest in the first-postfertilization year. Global Change Biol 17: 1442-1460.
http://dx.doi.org/10.111 l/i.l365-2486.2010.02298.x.
Chen, CY; Stemberger, RS: Kamman, NC: Mayes, BM; Folt, CL. (2005a). Patterns of Hg
bioaccumulation and transfer in aquatic food webs across multi-lake studies in the
Northeast US. Ecotoxicology 14: 135-147. http://dx.doi.org/10.1007/slQ646-004-6265-v.
Chen, D; Lan, Z; Bai, X: Grace, JB; Bai, Y. (2013). Evidence that acidification-induced declines
in plant diversity and productivity are mediated by changes in below-ground
communities and soil properties in a semi-arid steppe. J Ecol 101: 1322-1334.
http://dx.doi.org/10.1111/1365-2745.12119.
Chen, D; Lan, Z; Hu, S: Bai, Y. (2015). Effects of nitrogen enrichment on belowground
communities in grassland: Relative role of soil nitrogen availability vs. soil acidification.
Soil Biol Biochem 89: 99-108. http://dx.doi.Org/10.1016/i.soilbio.2015.06.028.
Chen, L; Driscoll, CT. (2005). A two-layer model to simulate variations in surface water
chemistry draining a northern forest watershed. Water Resour Res 41: W09425.
http://dx.doi. org/10.1029/2004WR003 625.
Chen, S: Bai, Y; Zhang, L; Han, X. (2005b). Comparing physiological responses of two
dominant grass species to nitrogen addition in Xilin River Basin of China. Environ Exp
Bot 53: 65-75. http://dx.doi.Org/10.1016/i.envexpbot.2004.03.002.
Chen. Y. -L: Zhang. X: Ye. J. -S: Han. H, -Y: Wan. S. -O: Chen. B. -D. (20141 Six-year
fertilization modifies the biodiversity of arbuscular mycorrhizal fungi in a temperate
steppe in Inner Mongolia. Soil Biol Biochem 69: 371-381.
http://dx.doi.Org/10.1016/i.soilbio.2013.l 1.020.
June 2018
R-40
DRAFT: Do Not Cite or Quote

-------
Chen. YL; Ding. J. inZhi; Peng. Y. unF; Li. F. ei; Yang. G, uiB; Liu. L, i; Qin, S, huQi; Fang. K,
ai; Yang. YH. e. (2016). Patterns and drivers of soil microbial communities in Tibetan
alpine and global terrestrial ecosystems. Journal of Biogeography 43: 2027-2039.
http://dx.doi.org/10. Ill 1/ibi. 12806.
Chen. ZJ: Geng. SC: Zhang. JH: Setala. H: Gu. Y; Wang. F; Zhang. X: Wang. XX: Han. SJ.
(2017). Addition of nitrogen enhances stability of soil organic matter in a temperate
forest. Eur J Soil Sci 68: 189-199. http://dx.doi.org/10.Ill 1/ejss. 12404.
Cheng. X: Luo. Y; Xu. Q: Lin. G: Zhang. 0; Chen. J: Li. B. (2010). Seasonal variation in CH4
emission and its 13C-isotopic signature from Spartina alterniflora and Scirpus mariqueter
soils in an estuarine wetland. Plant Soil 327: 85-94. http://dx.doi.org/10.1007/sl 1104-
009-0033-y.
Chestnut. LG: Mills. DM. (2005). A fresh look at the benefits and costs of the US acid rain
program. J Environ Manage 77: 252-266.
http://dx.doi.Org/10.1016/i.ienvman.2005.05.014.
Chiwa. M; Sheppardb. LJ; Leith. I; Leeson. S. R.: Tang. Y; Cape. JN. (2016). Sphagnum can
'filter' N deposition, but effects on the plant and porewater depend on the N form. Sci
Total Environ 559: 113-120. http://dx.doi.Org/10.1016/i.scitotenv.2016.03.130.
Chmura. GL; Kellman. L; van Ardenne. L. ee: Guntenspergen. GR. (2016). Greenhouse Gas
Fluxes from Salt Marshes Exposed to Chronic Nutrient Enrichment. PLoS ONE 11:
e0149937. http://dx.doi.org/10.1371/iournal.pone.0149937.
Cho, Y; Driscoll, CT; Blum, JD. (2009). The effects of a whole-watershed calcium addition on
the chemistry of stream storm events at the Hubbard Brook Experimental Forest in NH,
USA. Sci Total Environ 407: 5392-5401.
http://dx.doi.Org/10.1016/i.scitotenv.2009.06.030.
Cho, Y; Driscoll, CT; Johnson, CE; Blum, JD; Fahey, TJ. (2012). Watershed-level responses to
calcium silicate treatment in a northern hardwood forest. Ecosystems 15: 416-434.
http://dx.doi.org/10.1007/slQ021-012-9518-2.
Chodak, M; Golebiewski, M; Morawska-Ploskonka, J: Kuduk, K; Niklinska, M. (2013).
Diversity of microorganisms from forest soils differently polluted with heavy metals.
Appl Soil Ecol 64: 7-14. http://dx.doi.Org/10.1016/i.apsoil.2012.l 1.004.
Chu, C: Bartlett, M; Wang, Y; He, F; Weiner, J: Chave, J: Sack, L. (2016). Does climate directly
influence NPP globally? Global Change Biol 22: 12-24.
http://dx.doi.org/10. Ill 1/gcb. 13079.
Churchland, C: Mavo-Bruinsma, L; Ronson, A: Grogan, P. (2010). Soil microbial and plant
community responses to single large carbon and nitrogen additions in low arctic tundra.
Plant Soil 334: 409-421. http://dx.doi.org/10.1007/slllQ4-010-0392-4.
Churkina, G: Zaehle, S: Hughes, J: Viovv, N: Chen, Y; Jung, M; Heumann, BW: Ramankuttv,
N: Heimann, M; Jones, C. (2010). Interactions between nitrogen deposition, land cover
conversion, and climate change determine the contemporary carbon balance of Europe.
Biogeosciences 7: 2749-2764. http://dx.doi.org/10.5194/bg-7-2749-2010.
June 2018
R-41
DRAFT: Do Not Cite or Quote

-------
Cione. NK; Padgett PE; Allen. EB. (2002). Restoration of a native shrubland impacted by exotic
grasses, frequent fire and nitrogen deposition in southern California. Restor Ecol 10: 376-
384. http://dx.doi.Org/10.1046/i.1526-100X.2002.02038.x.
Cira. EK; Paerl, HW: Wetz. MS. (2016). Effects of Nitrogen Availability and Form on
Phytoplankton Growth in a Eutrophied Estuary (Neuse River Estuary, NC, USA). PLoS
ONE 11: e0160663. http://dx.doi.org/10.1371/iournal.pone.0160663.
Citterio. S: Prato. N: Fumagalli. P; Aina. R: Massa. N: Santagostino. A: Sgorbati. S: Bert a. G.
(2005). The arbuscular mycorrhizal fungus Glomus mosseae induces growth and metal
accumulation changes in Cannabis sativa L. Chemosphere 59: 21-29.
http://dx.doi. org/10.1016/i. chemosphere.2004.10.009.
Civerolo. K; Hogrefe. C: Zalewsky. E: Hao. W: Sistla. G: Lynn. B: Rosenzweig. C: Kinney. PL.
(2010). Evaluation of an 18-year CMAQ simulation: Seasonal variations and long-term
temporal changes in sulfate and nitrate. Atmos Environ 44: 3745-3752.
http://dx.doi.Org/10.1016/i.atmosenv.2010.06.056.
Cizdziel. JV: Dempsev. S: Halbrook. RS. (2013). Preliminary evaluation of the use of homing
pigeons as biomonitors of mercury in urban areas of the USA and China. Bull Environ
Contam Toxicol 90: 302-307. http://dx.doi.org/10.1007/sQ0128-012-0918-v.
Clarisse. L; Shephard. MW: Dentener. F; Hurtmans. D; Cadv-Pereira. K; Karagulian. F; Van
Damme. M; Clerbaux. C: Coheur. P. -F. (2010). Satellite monitoring of ammonia: A case
study of the San Joaquin Valley. J Geophys Res Atmos 115: D13302.
http://dx.doi.org/10.1029/2009JDQ13291.
Clark. CM: Bell. MP: Boyd. JW: Compton. JE; Davidson. EA; Davis. C: Fenn. ME: Geiser. L;
Jones. L; Blett. TF. (2017). Nitrogen-induced terrestrial eutrophication: Cascading effects
and impacts on ecosystem services. Ecosphere 8: e01877.
http://dx.doi.org/10.1002/ecs2.1877.
Clark. CM: Cleland. EE: Collins. SL; Fargione. JE: Gough. L; Gross. KL; Pennings. SC: Suding.
KN: Grace. JB. (2007). Environmental and plant community determinants of species loss
following nitrogen enrichment. Ecol Lett 10: 596-607. http://dx.doi.Org/10.l 111/i. 1461-
0248.2007.01053.x.
Clark. CM: Morefield. PE: Gilliam. FS: Pardo. LH. (2013). Estimated losses of plant
biodiversity in the United States from historical N deposition (1985-2010). Ecology 94:
1441-1448. http://dx.doi.Org/10.1890/12-2016.l.
Clark, CM; Phelan, J: Doraiswamy, P; Buckley, J: Cajka, JC; Dennis, RL; Lynch, J: Nolte, CG;
Spero, TL. (2018). Atmospheric deposition and exceedances of critical loads from 1800-
2025 for the conterminous United States. Ecol Appl. http://dx.doi.org/10.1002/eap.1703.
Clark, CM: Tilman, D. (2008). Loss of plant species after chronic low-level nitrogen deposition
to prairie grasslands. Nature 451: 712-715. http://dx.doi.org/10.1038/nature06503.
Clarke, JF: Edgerton, ES: Martin, BE. (1997). Dry deposition calculations for the clean air status
and trends network. Atmos Environ 31: 3667-3678. http://dx.doi.org/10.lQ16/S1352-
2310(97^)00141-6.
June 2018
R-42
DRAFT: Do Not Cite or Quote

-------
Clarke. L. (2013). Rocky Mountain National Park Initiative air quality control commission
update February 21, 2013.
Clavden. MG; Lescord. GL; Kidd. KA; Wang. X; Muir. DC; O'Driscoll. NJ. (2017). Using sulfur
stable isotopes to assess mercury bioaccumulation and biomagnification in temperate lake
food webs. Environ Toxicol Chem 36: 661-670. http://dx.doi.org/10.10Q2/etc.3615.
Cleavitt. NL; Battles. JJ; Fahev. TJ; Blum. JD. (2014). Determinants of survival over 7 years for
a natural cohort of sugar maple seedlings in a northern hardwood forest. Can J For Res
44: 1112-1121. http://dx.doi.org/10.1139/cifr-2014-0177.
Cleavitt. NL: Ewing. HA: Weathers. KC: Lindsev. AM. (201 la). Acidic atmospheric deposition
interacts with tree type and impacts the cryptogamic epiphytes in Acadia National Park,
Maine, USA. Bryologist 114: 570-582. http://dx.doi.Org/10.1639/0007-2745-114.3.570.
Cleavitt. NL: Fahev. TJ: Battles. JJ. (201 lb). Regeneration ecology of sugar maple (Acer
saccharum): Seedling survival in relation to nutrition, site factors, and damage by insects
and pathogens. Can J For Res 41: 235-244. http://dx.doi.org/10.1139/X10-210.
Cleavitt. NL; Hinds. JW; Poirot. RL; Geiser. LH; Dibble. AC; Leon. B; Perron. R; Pardo. LH.
(2015). Epiphytic macrolichen communities correspond to patterns of sulfur and nitrogen
deposition in the northeastern United States. Bryologist 118: 304-324.
http://dx.doi.Org/10.1639/0007-2745-118.3.304.
Cleckner. LB; Gilmour. CC; Hurley. JP; Krabbenhoft. DP. (1999). Mercury methylation in
periphyton of the Florida Everglades. Limnol Oceanogr 44: 1815-1825.
http://dx.doi.Org/10.4319/lo.1999.44.7.1815.
Clemens. S. (2006). Toxic metal accumulation, responses to exposure and mechanisms of
tolerance in plants [Review], Biochimie 88: 1707-1719.
http://dx.doi.Org/10.1016/i.biochi.2006.07.003.
Clough. WS. (1975). The deposition of particles on moss and grass surfaces. Atmos Environ
(1967)9: 1113-1119. http://dx.doi.org/10.1016/0004-6981(75^)90187-0.
Clow. DW; Nanus. L; Huggett. B. (2010). Use of regression-based models to map sensitivity of
aquatic resources to atmospheric deposition in Yosemite National Park, USA. Water
ResourRes46: 1-14. http://dx.doi.org/10.1029/2009wr008316.
Clow. DW; Roop. HA; Nanus. L; Fenn. ME; Sexstone. GA. (2015). Spatial patterns of
atmospheric deposition of nitrogen and sulfur using ion-exchange resin collectors in
Rocky Mountain National Park, USA. Atmos Environ 101: 149-157.
http://dx.doi.Org/10.1016/i.atmosenv.2014.l 1.027.
Clow. DW; Sickman. JO; Striegl. RG; Krabbenhoft. DP; Elliot. JG; Dornblaser. M; Roth. DA;
Campbell. DH. (2003). Changes in the chemistry of lakes and precipitation in high-
elevation national parks in the western United States, 1985-1999. Water Resour Res 39:
4-1 - 4-13. http://dx.doi.org/10.1029/2002WR0Q1533.
June 2018
R-43
DRAFT: Do Not Cite or Quote

-------
Clow. DW; Sueker. JK. (2000). Relations between basin characteristics and stream water
chemistry in alpine/subalpine basins in Rocky Mountain National Park, Colorado. Water
ResourRes 36: 49-61. http://dx.doi.org/10.1029/1999WR90Q294.
Clow. DW: Williams. MW; Schuster. PF. (2016). Increasing aeolian dust deposition to
snowpacks in the Rocky Mountains inferred from snowpack, wet deposition, and aerosol
chemistry. Atmos Environ 146: 183-194.
http://dx.doi.ore/10.1016/i.atmosenv.2016.06.076.
Coelho. JP; Rosa. M; Pereira. E; Duarte. A: Pardal. MA. (2006). Pattern and annual rates of
Scrobicularia plana mercury bioaccumulation in a human induced mercury gradient (Ria
de Aveiro, Portugal). Estuar Coast Mar Sci 69: 629-635.
http://dx.doi.ore/10.1016/j.ecss.2006.05.027.
Cole. DW: Rapp. M. (1981). Elemental cycling in forest ecosystems. In DE Reiche (Ed.),
Dynamic properties of forest ecosystems (pp. 341-409). Cambidge, UK: Cambridge
University Press.
http://www.cambridee.ore/cataloeue/cataloeue.asp?isbn=9780521112208.
Cole. JJ; Prairie. YT; Caraco. NF; McDowell. WH; Tranvik. LJ; Strieel. RG; Duarte. CM:
Kortelainen. P; Downine. JA; Middelbure. JJ: Melack. J. (2007). Plumbing the global
carbon cycle: Integrating inland waters into the terrestrial carbon budget. Ecosystems 10:
172-185. http://dx.doi.ore/10.1007/sl0021-006-9013-8.
Collins. M; Knutti. R; Arblaster. J: Dufresne. JL; Fichefet. T; Friedlinestein. P; Gao. X:
Gutowski. J. Jr; Johns. T; Krinner. G: Shonewe. M; Tebaldi. C: Weaver. AJ; Wehner. M.
(2014). Long-term climate change: Projections, commitments and irreversibility. In TF
Stocker; D Qin; G Plattner; MMB Tignor; SK Allen; J Boschung; A Mauels; Y Xia; V
Bex; PM Midgley (Eds.), Climate Change 2013 The Physical ScienceBasis (pp. 1029-
1136). New York, NY: Cambridge University Press.
http://dx.doi.ore/10.1017/CBQ9781107415324.024.
Collins, RN; Merrineton, G; McLauehlin, MJ; Morel, JL. (2003). Organic ligand and pH effects
on isotopically exchangeable cadmium in polluted soils. Soil Sci Soc Am J 67: 112-121.
Collins, SL; Knapp, AK; Briees, JM; Blair, JM; Steinauer, EM. (1998). Modulation of diversity
by grazing and mowing in native tallgrass prairie. Science 280: 745-747.
http://dx.doi.ore/10.1126/science.280.5364.745.
Collins, SL: Ladwie, LM; Petrie, MP: Jones, SK: Mulhouse, JM: Thibault, JR: Pockman, WT.
(2017). Press-pulse interactions: effects of warming, N deposition, altered winter
precipitation, and fire on desert grassland community structure and dynamics. Global
Change Biol 23: 1095-1108. http://dx.doi.ore/10. Ill 1/ecb. 13493.
Compton, JE: Harrison, JA: Dennis, RL: Greaver, TL: Hill, BH: Jordan, SJ: Walker, H:
Campbell, HV. (2011). Ecosystem services altered by human changes in the nitrogen
cycle: A new perspective for US decision making [Review], Ecol Lett 14: 804-815.
http://dx.doi.ore/10.ll 11/i. 1461-0248.2011.01631.x.
June 2018
R-44
DRAFT: Do Not Cite or Quote

-------
Compton. JE; Hill B; Dennis. R. (2013). Connecting nitrogen deposition and ecosystem
services. In RA Pielke, Sr.; TR Seastedt; K Suding (Eds.), Reference Module in Earth
Systems and Environmental Sciences - Climate Vulnerability (pp. 2333). Burlington,
MA: Academic Press. http://dx.doi.org/10.1016/B978-0-12-384703-4.004Q3-2.
Compton. JE: Watrud. LS: Porteous. A: DeGrood. S. (2004). Response of soil microbial biomass
and community composition to chronic nitrogen additions at Harvard forest. For Ecol
Manage 196: 143-158. http://dx.doi.Org/10.1016/i.foreco.2004.03.017.
Conant. RT; Ryan. MG: Agren. GI; Birge. HE: Davidson. EA; Eliasson. PE; Evans. SE; Frev.
SD; Giardina. CP: Hopkins. FM; Hyvonen. R; Kirschbaum. MUF; Lavallee. JM; Leifeld.
J: Parton. WJ: Steinweg. JM: Wallenstein. MP: Wetterstedt. JAM: Bradford. MA.
(2011). Temperature and soil organic matter decomposition rates - Synthesis of current
knowledge and away forward. Global Change Biol 17: 3392-3404.
http://dx.doi.org/10.1111/i. 1365-2486.2011,02496.x.
Concilio. AL; Loik. ME. (2013). Elevated nitrogen effects on Bromus tectorum dominance and
native plant diversity in an arid montane ecosystem. Appl Veg Sci 16: 598-609.
http://dx.doi.org/10. Ill 1/avsc. 12029.
Concilio. AL: Nippert. JB; Ehrenfeucht. S: Cherwin. K; Seastedt. TR. (2016). Imposing
antecedent global change conditions rapidly alters plant community composition in a
mixed-grass prairie. Oecologia 182: 899-911. http://dx.doi.org/10.1007/sQ0442-016-
3684-4.
Conlev. DJ: Paerl. HW: Howarth. RW: Boesch. DF; Seitzinger. SP; Havens. KE; Lancelot. C:
Likens. GE. (2009). Controlling eutrophication: Nitrogen and phosphorus. Science 323:
1014-1015. http://dx.doi.org/10.1126/science.l 167755.
Cook. RB; Elwood. JW: Turner. RR; Bogle. MA: Mulholland. PJ: Palumbo. AY. (1994). Acid-
base chemistry of high-elevation streams in the Great Smoky Mountains. Water Air Soil
Pollut 72: 331-356. http://dx.doi.org/10.1007/BF01257133.
Coolev. SR; Donev. SC. (2009). Anticipating ocean acidification's economic consequences for
commercial fisheries. Environ Res Lett 4: Article #024007.
http://dx.doi.Org/10.1088/1748-9326/4/2/024007.
Coolev. SR: Rheuban. JE: Hart. PR: Luu. V: Glover. DM: Hare. JA; Donev. SC. (2015). An
Integrated Assessment Model for Helping the United States Sea Scallop (Placopecten
magellanicus) Fishery Plan Ahead for Ocean Acidification and Warming. PLoS ONE 10:
e0124145. http://dx.doi.org/10.1371/iournal.pone.0124145.
Cooper. AB. (1990). Nitrate depletion in the riparian zone and stream channel of a small
headwater catchment. Hydrobiologia 202: 13-26. http://dx.doi.org/10.1007/bf000278Q9.
Cooper. E: Neil. K: Gaiewski. K. (2016). Spatial and temporal cladoceran community responses
to environmental change and anthropogenic impacts in southwestern Quebec. Ecoscience
23: 97-112. http://dx.doi.org/10.1080/11956860.2Q17.1286895.
Cooper. SR: Brush. GS. (1993). A 2,500-year history of anoxia and eutrophication in
Chesapeake Bay. Estuaries 16: 617-626. http://dx.doi.org/10.2307/1352799.
June 2018
R-45
DRAFT: Do Not Cite or Quote

-------
Corkidi. L; Rowland. PL; Johnson. NC; Allen. EB. (2002). Nitrogen fertilization alters the
functioning of arbuscular mycorrhizas at two semiarid grasslands. Plant Soil 240: 299-
310. http://dx.doi.Org/10.1023/A:1015792204633.
Cornell. S. (201 la). Atmospheric nitrogen deposition: Revisiting the question of the invisible
organic fraction. Procedia Environ Sci 6: 96-103.
http://dx.doi.Org/10.1016/i.proenv.2011.05.010.
Cornell. SE. (201 lb). Atmospheric nitrogen deposition: Revisiting the question of the
importance of the organic component. Environ Pollut 159: 2214-2222.
http://dx.doi.Org/10.1016/i.envpol.2010.l 1.014.
Cornwell. JC: Glibert. PM; Owens. MS. (2014). Nutrient Fluxes from Sediments in the San
Francisco Bay Delta. Estuaries Coasts 37: 1120-1133. http://dx.doi.org/10.10Q7/sl2237-
013-9755-4.
Corrales, J; Naja, GM; Dziuba, C; Rivero, RG; Orem, W. (2011). Sulfate threshold target to
control methylmercury levels in wetland ecosystems. Sci Total Environ 409: 2156-2162.
http://dx.doi.Org/10.1016/i.scitotenv.2011.02.030.
Correia. RRS: Miranda. MR: Guimaraes. JRD. (2012). Mercury methylation and the microbial
consortium in periphyton of tropical macrophytes: Effect of different inhibitors. Environ
Res 112: 86-91. http://dx.doi.Org/10.1016/i.envres.2011.11.002.
Cosby. BJ: Hornberger. GM: Galloway. JN; Wright. RF. (1985a). Modeling the effects of acid
deposition: Assessment of a lumped parameter model of soil water and streamwater
chemistry. Water Resour Res 21: 51-63. http://dx.doi.org/10.1029/WR021i001pQ0051.
Cosby. BJ: Hornberger. GM: Galloway. JN: Wright. RF. (1985b). Time scales of catchment
acidification: A quantitative model for estimating freshwater acidification. Environ Sci
Technol 19: 1144-1149. http://dx.doi.org/10.1021/esQ0142a001.
Cosby. BJ: Webb. JR; Galloway. JN: Devinev. FA. (2006). Acidic deposition impacts on natural
resources in Shenandoah National Park. (NPS/NER/NRTR--2006/066). Philadelphia, PA:
U.S. Department of the Interior, National Park Service, Northeast Region.
Cosby. BJ: Wright. RF: Hornberger. GM: Galloway. JN. (1985c). Modeling the effects of acid
deposition: Estimation of long-term water quality responses in a small forested
catchment. Water Resour Res 21: 1591-1601.
http://dx.doi.org/10.1029/WR021i011pQ1591.
Cosme. N: Jones. MC: Cheung. WWL: Larsen. HF. (2017). Spatial differentiation of marine
eutrophication damage indicators based on species density. Ecol Indicat 73: 676-685.
http://dx.doi.Org/10.1016/i.ecolind.2016.10.026.
Costanza, R; dArge, R; de Groot, R; Farber, S; Grasso, M; Hannon, B; Limburg, K; Naeem, S;
O'Neill. RV: Paruelo. J: Raskin. RG: Sutton. P: van den Belt. M. (1997). The value of the
world's ecosystem services and natural capital. Nature 387: 253-259.
http://dx.doi.org/10.1038/387253a0.
June 2018
R-46
DRAFT: Do Not Cite or Quote

-------
Costanza. R; De Groot R: Braat L; Kubiszewski. I; Fioramonti. L; Sutton. P; Farber. S; Grasso.
M. (2017). Twenty years of ecosystem services: How far have we come and how far do
we still need to go? Ecosyst Serv 28: 1-16.
http://dx.doi.Org/10.1016/i.ecoser.2017.09.008.
Costello. CT; Ken worth v. WJ. (2011). Twelve-year mapping and change analysis of Eelgrass
(Zostera marina) areal abundance in Massachusetts (USA) identifies statewide declines.
Estuaries Coasts 34: 232-242. http://dx.doi.org/10.1007/sl2237-01Q-9371-5.
Couture. S: Houle, D; Gagnon. C. (2012). Increases of dissolved organic carbon in temperate and
boreal lakes in Quebec, Canada. Environ Sci PollutRes Int 19: 361-371.
http://dx.doi.org/10.1007/sll356-011-Q565-6.
Cowardin. LM; Carter. V: Golet. FC: LaRoe. ET. (1979). Classification of wetland and
deepwater habitats of the United States. (FWS/OBS-79/31). Washington, DC: U.S.
Department of the Interior, Fish and Wildlife Service.
Cox. F; Barsoum. N: Lilleskov. EA; Bidartondo. MI. (2010). Nitrogen availability is a primary
determinant of conifer mycorrhizas across complex environmental gradients. Ecol Lett
13: 1103-1113. http://dx.doi.org/10.1111/i. 1461-0248.2010.01494.X.
Cox, RD; Preston, KL; Johnson, RF; Minnich, RA; Allen, EB. (2014). Influence of landscape-
scale variables on vegetation conversion to exotic annual grassland in southern
California, USA. Global Ecology and Conservation 2: 190-203.
http://dx.doi.Org/10.1016/i.gecco.2014.09.008.
Creed, IF; Morrison, PL; Nicholas, NS. (2004). Is coarse woody debris a net sink or source of
nitrogen in the red spruce - Fraser fir forest of the southern Appalachians, USA? Can J
For Res 34: 716-727. http://dx.doi.org/10.1139/X03-211.
Creswell, JE; Kerr, SC: Meyer, MH; Babiarz, CL; Shafer, MM: Armstrong, DE: Roden, EE.
(2008). Factors controlling temporal and spatial distribution of total mercury and
methylmercury in hyporheic sediments of the Allequash Creek wetland, northern
Wisconsin. J Geophys Res 113: 1-9. http://dx.doi.org/10.1029/2008JG00Q742.
Creswell, JE: Shafer, MM: Babiarz, CL: Tan, SZ; Musinsky, AL; Schott, TH; Roden, EE:
Armstrong, DE. (2017). Biogeochemical controls on mercury methylation in the
Allequash Creek wetland. Environ Sci Pollut Res Int 24: 15325-15339.
http://dx.doi. org/10.1007/sl 1356-017-9094-2.
Cronan, CS: Grigal, DF. (1995). Use of calcium/aluminum ratios as indicators of stress in forest
ecosystems. J Environ Qual 24: 209-226.
http://dx.doi.org/10.2134/ieal995.004724250024000200Q2x.
Cronan, CS: Reiners, WA; Reynolds, RCJ: Lang, GE. (1978). Forest floor leaching:
Contributions from mineral, organic, and carbonic acids in New Hampshire subalpine
forests. Science 200: 309-311. http://dx.doi.Org/10.l 126/science.200.4339.309-a.
Cronan, CS: Schofield, CL. (1990). Relationships between aqueous aluminum and acidic
deposition in forested watersheds of North America and northern Europe. Environ Sci
Technol 24: 1100-1105. http://dx.doi.org/10.1021/esQ0077a022.
June 2018
R-47
DRAFT: Do Not Cite or Quote

-------
Croteau. MN; Luoma. SN; Stewart. AR. (2005). Trophic transfer of metals along freshwater food
webs: Evidence of cadmium biomagnification in nature. Limnol Oceanogr 50: 1511-
1519.
Crowe. SA; Canfield. DE; Mucci. A: Sundbv. B: Maranger, R. (2012). Anammox, denitrification
and fixed-nitrogen removal in sediments from the Lower St. Lawrence Estuary.
Biogeosciences 9: 4309-4321. http://dx.doi.org/10.5194/bg-9-4309-2Q12.
Crumley. KM: Teece. MA: Crandall. JB; Sauer. AK; Driscoll. CT. (2016). Effects of nitrogen
deposition on nitrogen acquisition by Sarracenia purpurea in the Adirondack Mountains,
New York, USA. Torrey Botanical Society Journal 143: 8-20.
http://dx.doi.org/10.3159/TORREY-D-14-0008Q.
Csavina. J: Roberti. JA; Taylor. JR; Loescher. HW. (2017). Traceable measurements and
calibration: a primer on uncertainty analysis. Ecosphere 8.
http://dx.doi.org/10.1002/ecs2.1683.
Curry. JA: Webster. PJ. (2011). CLIMATE SCIENCE AND THE UNCERTAINTY
MONSTER. Bull Am Meteorol Soc 92: 1667-+.
http://dx.doi.Org/10.1175/2011BAMS3139.l.
Curtis, CJ; Heaton, THE; Simpson, GL; Evans, CD: Shilland, J: Turner, S. (2012). Dominance of
biologically produced nitrate in upland waters of Great Britain indicated by stable
isotopes. Biogeochemistry 111: 535-554. http://dx.doi.org/10.1007/slQ533-011-9686-8.
Curtis, CJ: Kaiser, J, an: Marca, A: Anderson, NJ: Simpson, G: Jones, V: Whiteford, E. (2018).
Spatial variations in snowpack chemistry, isotopic composition of N03- and nitrogen
deposition from the ice sheet margin to the coast of western Greenland. Biogeosciences
15: 529-550. http://dx.doi.org/10.5194/bg-15-529-2018.
Cusack, DF; Torn, MS: McDowell, WH: Silver, WL. (2010). The response of heterotrophic
activity and carbon cycling to nitrogen additions and warming in two tropical soils.
Global Change Biol 16: 2555-2572. http://dx.doi.Org/10.llll/i.1365-2486.2009.02131.x.
D'Andrea, AF: DeWitt, TH. (2009). Geochemical ecosystem engineering by the mud shrimp
Upogebia pugettensis (Crustacea: Thalassinidae) in Yaquina Bay, Oregon: Density-
dependent effects on organic matter remineralization and nutrient cycling. Limnol
Oceanogr 54: 1911-1932. http://dx.doi.Org/10.4319/lo.2009.54.6.1911.
D'Angelo, C: Wiedenmann, J. (2014). Impacts of nutrient enrichment on coral reefs: New
perspectives and implications for coastal management and reef survival. Current Opinion
in Environmental Sustainability 7: 82-93. http://dx.doi.Org/10.1016/i.cosust.2013.l 1.029.
D'Elia, CF: Sanders, JG: Boynton, WR. (1986). Nutrient enrichment studies in a coastal plain
estuary: phytoplankton growth in large-scale, continuous cultures. Can J Fish Aquat Sci
43: 397-406.
Da Silva, LC: Oliva, MA: Azevedo, AA: De Arauio, JM. (2006). Responses of restinga plant
species to pollution from an iron pelletization factory. Water Air Soil Pollut 175: 241-
256. http://dx.doi.org/10.1007/sll270-0Q6-9135-9.
June 2018
R-48
DRAFT: Do Not Cite or Quote

-------
Daebeler. A; Bodelier. PLE; Hefting. MM; Rutting. T; Jia. Z; Laanbroek. HJ. (2017). Soil
warming and fertilization altered rates of nitrogen transformation processes and selected
for adapted ammonia-oxidizing archaea in sub-arctic grassland soil. Soil Biol Biochem
107: 114-124. http://dx.doi.Org/10.1016/i.soilbio.2016.12.013.
Daggett. CT; Saros. JE; Lafrancois. BM; Simon. KS: Amirbahman. A. (2015). Effects of
increased concentrations of inorganic nitrogen and dissolved organic matter on
phytoplankton in boreal lakes with differing nutrient limitation patterns. Aquat Sci 77:
511-521. http://dx.doi.org/10.1007/s00027-015-Q396-5.
Dahlman. L; Nasholm. T; Palmqvist. K. (2002). Growth, nitrogen uptake, and resource
allocation in the two tripartite lichens Nephroma arcticum and Peltigera aphthosa during
nitrogen stress. NewPhytol 153: 307-315. http://dx.doi.org/10.1046/j.0028-
646X.2001.00321.x.
Daily. GC. (1997). Introduction: What are ecosystem services? In GC Daily (Ed.), Societal
Dependence on Natural Ecosystems (pp. 1-10). Washington, DC: Island Press.
Dale. AW: Sommer. S: Bohlen. L; Treude. T; Bertics. VJ: Bange. HW; Pfannkuche. O: Schorp.
T; Mattsdotter. M; Wallmann. K. (2011). Rates and regulation of nitrogen cycling in
seasonally hypoxic sediments during winter (Boknis Eck, SW Baltic Sea): Sensitivity to
environmental variables. Estuar Coast Shelf Sci 95: 14-28.
http://dx.doi.Org/10.1016/i.ecss.2011.05.016.
Dale. VH; Kling. CL; Meyer. JL; Sanders. J: Stallworth. H; Armitage. T; Wangsness. D;
Bianchi. T; Blumberg. A: Boynton. W: Conlev. DJ: Crumpton. W: David. M; Gilbert. D;
Howarth. RW: Lowrance. R; Mankin. K; Opaluch. J: Paerl. H; Reckhow. K; Sharplev.
AN: Simpson, TW; Snyder, CS; Wright, D; Dale, VH; Kling, CL; Meyer, JL; Sanders, J:
Stallworth, H; Armitage, T; Wangsness, D; Bianchi, T; Blumberg, A: Boynton, W:
Conlev, DJ: Crumpton, W: David, M; Gilbert, D; Howarth, RW: Lowrance, R; Mankin,
K; Opaluch, J: Paerl, H; Reckhow, K; Sharplev, AN: Simpson, TW: Snyder, CS: Wright,
D. (2010). Hypoxia in the Northern Gulf of Mexico a brief overview introduction. In
Hypoxia in the Northern Gulf of Mexico. New Tork, NY: Springer New York.
http://dx.doi.org/10.1007/978-0-387-89686-l 1.
Dallas, T; Drake, JM. (2014). Nitrate enrichment alters a Daphnia-microparasite interaction
through multiple pathways. Ecol Evol 4: 243-250. http://dx.doi.org/10.1002/ece3.925.
Damashek, J; Casciotti, KL; Francis, CA. (2016). Variable Nitrification Rates Across
Environmental Gradients in Turbid, Nutrient-Rich Estuary Waters of San Francisco Bay.
Estuaries Coasts 39: 1050-1071. http://dx.doi.org/10.1007/sl2237-016-0Q71-7.
Damashek, J: Smith, JM; Mosier, AC: Francis, CA. (2015). Benthic ammonia oxidizers differ in
community structure and biogeochemical potential across a riverine delta. FMICB 5: 743.
http://dx.doi.org/10.3389/fmicb.2014.00743.
Dana, M. (1980). S02 versus sulfate wet deposition in the eastern United States. J Geophys Res
85: 4475-4480.
Darby, FA; Turner, RE. (2008). Below- and aboveground biomass of Spartina alterniflora:
Response to nutrient addition in a Louisiana salt marsh. Estuaries Coasts 31: 326-334.
http://dx.doi.org/10.1007/sl2237-008-9Q37-8.
June 2018
R-49
DRAFT: Do Not Cite or Quote

-------
Daresta. BE; Italiano. F; Gennaro. GD; Trotta. M; Tutino. M; Veronico. P. (2014). Atmospheric
particulate matter (PM) effect on the growth of Solanum lycopersicum cv. Roma plants.
Chemosphere 119C: 37-42. http://dx.doi.Org/10.1016/i.chemosphere.2014.05.054.
Darveau. M; Huot J: Belanger, L. (1998). Riparian forest strips as habitat for snowshoe hare in a
boreal balsam fir forest. Can J For Res 28: 1494-1500. http://dx.doi.Org/10.l 139/x98-l 10.
Darveau. M; Labbe, P; Beauchesne, P; Belanger, L; Huot J. (2001). The use of riparian forest
strips by small mammals in a boreal balsam fir forest. For Ecol Manage 143: 95-104.
Das. B; Vinebrooke. RD; Sanchez-Azofeifa. A: Rivard, B; Wolfe. AP. (2005). Inferring
sedimentary chlorophyll concentrations with reflectance spectroscopy: A novel approach
to reconstructing historical changes in the trophic status of mountain lakes. Can J Fish
Aquat Sci 62: 1067-1078. http://dx.doi.org/10.1139/F05-016.
Dastis. JON: Perry. AM. (2016). The impact of regional landscape context on local maladaptive
trait divergence: a field test using freshwater copepod acid tolerance. Evolutionary
Ecology 30: 841-859. http://dx.doi.org/10.1007/slQ682-016-9853-4.
Davev. E; Wigand. C: Johnson. R; Sundberg. K; Morris. J: Roman. CT. (2011). Use of computed
tomography imaging for quantifying coarse roots, rhizomes, peat, and particle densities in
marsh soils. Ecol Appl 21: 2156-2171. http://dx.doi.org/10.1890/10-2037.1.
David. MB: Lawrence. GB. (1996). Soil and soil solution chemistry under red spruce stands
across the northeastern United States. Soil Sci 161: 314-328.
Davidson. CI: Miller. JM; Pleskow. MA. (1982). The influence of surface structure on predicted
particle dry deposition to natural grass canopies. Water Air Soil Pollut 18: 25-43.
http://dx.doi.org/10.1007/BFQ2419401.
Davidson, K; Gowen, RJ; Tett, P; Bresnan, E; Harrison, PJ; McKinney, A; Milligan, S; Mills,
DK; Silke, J: Crooks, A, -M. (2012). Harmful algal blooms: How strong is the evidence
that nutrient ratios and forms influence their occurrence? Estuar Coast Shelf Sci 115:
399-413. http://dx.doi.Org/10.1016/i.ecss.2012.09.019.
Davidson, TA; Saver, CD: Langdon, PG; Burgess, AMY; Jackson, M. (2010a). Inferring past
zooplanktivorous fish and macrophyte density in a shallow lake: application of a new
regression tree model. 55: 584-599. http://dx.doi.Org/10.l 111/i. 1365-2427.2009.02391.x.
Davidson, TA: Saver, CD: Perrow, M; Bramm, M; Jeppesen, E. (2010b). The simultaneous
inference of zooplanktivorous fish and macrophyte density from sub-fossil cladoceran
assemblages: a multivariate regression tree approach. Freshw Biol 55: 546-564.
http://dx.doi.org/10.1111/i. 1365-2427.2008.02124.x.
Davies, L; Bates, JW: Bell, JNB; James, PW: Purvis, OW. (2007). Diversity and sensitivity of
epiphytes to oxides of nitrogen in London. Environ Pollut 146: 299-310.
http://dx.doi.Org/10.1016/i.envpol.2006.03.023.
Davies, SP; Jackson, SK. (2006). The biological condition gradient: A descriptive model for
interpreting change in aquatic ecosystems. Ecol Appl 16: 1251-1266.
http://dx.doi.org/10.1890/1051-0761(2006)016ri251:TBCGAD12.0.CQ:2.
June 2018
R-50
DRAFT: Do Not Cite or Quote

-------
Davis. JE; Goldstein. RA. (1988). Simulated response of an acidic Adirondack Lake Watershed
to various liming mitigation strategies. Water Resour Res 24: 525.
http://dx.doi.org/10.1029AVR024i004pQ0525.
Davis. MA: Grime. JP; Thompson. K. (2000). Fluctuating resources in plant communities: A
general theory of invasibility. J Ecol 88: 528-534. http://dx.doi.org/10.1046/i. 1365-
2745.2000.00473.x.
Davis. TW: Bulleriahn. GS: Tuttle. T; Mckav. RM; Watson. SB. (2015). Effects of increasing
nitrogen and phosphorus concentrations on phytoplankton community growth and
toxicity during planktothrix blooms in Sandusky Bay, Lake Erie. Environ Sci Technol 49:
7197-7207. http://dx.doi.org/10.1021/acs.est.5b00799.
Davis. TW: Harke. MJ; Marcoval. MA: Goleski. J: Orano-Dawson. C: Berry. PL: Gobler. CJ.
(2010).	Effects of nitrogenous compounds and phosphorus on the growth of toxic and
non-toxic strains of Microcystis during cyanobacterial blooms. Aquat Microb Ecol 61:
149-162. http://dx.doi.org/10.3354/ame01445.
Dawson. JJC; Malcolm. IA; Middlemas. SJ: Tetzlaff. D; Soulsbv. C. (2009). Is the composition
of dissolved organic carbon changing in upland acidic streams? Environ Sci Technol 43:
7748-7753. http://dx.doi.org/10.1021/es901649b.
Davvani. S: Prasher. SO: Madani. A: Madramootoo. CA; Enright. P. (2013). Evaluation of
WARMF model for flow and nitrogen transport in an agricultural watershed under a cold
climate. Water Qual Res J Can 48: 400-413. http://dx.doi.org/10.2166/wqric.2013.034.
De Barba, D; Rossi, S; Deslauriers, A; Morin, H. (2016). Effects of soil warming and nitrogen
foliar applications on bud burst of black spruce. Trees Struct Funct 30: 87-97.
http://dx.doi.org/10.1007/s00468-015-1152-Q.
de Bellefeuille. S: Belanger. L; Huot. J: Cimon. A. (2001). Clear-cutting and regeneration
practices in Quebec boreal balsam fir forest: Effects on snowshoe hare. Can J For Res 31:
41-51. http://dx.doi.Org/10.l 139/x00-140.
de Sassi. C: Lewis. OT; Tylianakis. JM. (2012). Plant-mediated and nonadditive effects of two
global change drivers on an insect herbivore community. Ecology 93: 1892-1901.
http://dx.doi.Org/10.1890/l 1-1839.1.
De Schriiver. A: De Frenne. P; Ampoorter. E; Van Nevel. L; Demev. A: Wuvts. K; Verheven. K.
(2011).	Cumulative nitrogen input drives species loss in terrestrial ecosystems. Global
Ecology and Biogeography 20: 803-816. http://dx.doi.org/10. Ill 1/i .1466-
8238.2011.00652.x.
De Schriiver. A: Verheven. K; Mertens. J: Staelens. J: Wuvts. K; Muvs. B. (2008). Nitrogen
saturation and net ecosystem production. Nature 451: El.
http://dx.doi.org/10.1038/nature06578.
de Vries. W: Du. E; Butterbach-Bahl. K. (2014a). Short and long-term impacts of nitrogen
deposition on carbon sequestration by forest ecosystems. Current Opinion in
Environmental Sustainability 9-10: 90-104.
http://dx.doi.Org/10.1016/i.cosust.2014.09.001.
June 2018
R-51
DRAFT: Do Not Cite or Quote

-------
de Vries. W; Goodale. C; Erisman. JW; Hettelingh: J-P. (2014b). Impacts of nitrogen deposition
on ecosystem services in interaction with other nutrients, air pollutants and climate
change. 4: 387-396. http://dx.doi.org/10.1007/978-94-0Q7-7939-6 41.
de Vries. W; Posch, M. (2011). Modelling the impact of nitrogen deposition, climate change and
nutrient limitations on tree carbon sequestration in Europe for the period 1900-2050.
EnvironPollut 159: 2289-2299. http://dx.doi.Org/10.1016/i.envpol.2010.l 1.023.
de Vries. W: Posch, M; Reinds, GJ: Hettelingh. J. -P. (2014c). Quantification of impacts of
nitrogen deposition on forest ecosystem services in Europe. In Nitrogen Deposition,
Critical Loads and Biodiversity. Netherlands: Springer. http://dx.doi.org/10.10Q7/978-94-
007-7939-6 43.
De Vries. W: Reinds. GJ: Gundersen. P; Sterba. H. (2006). The impact of nitrogen deposition on
carbon sequestration in European forests and forest soils. Global Change Biol 12: 1151-
1173. http://dx.doi.Org/10.llll/i.1365-2486.2006.01151.x.
de Vries. W: Solberg. S: Dobbertin. M; Sterba. H; Laubhahn. D; Reinds. GJ: Nabuurs. G. -J:
Gundersen. P; Sutton. MA. (2008). Ecologically implausible carbon response?
[Comment], Nature 451: E1-E3. http://dx.doi.org/10.1038/nature06579.
De Vries, W; Wamelink, GWW; van Dobben, H; Kros, J: Reinds, GJ: Mol-Dijkstra, JP; Smart,
SM; Evans, CD: Rowe, EC: Belvazid, S: Sverdrup, HU; van Hinsberg, A: Posch, M;
Hettelingh, J, -P; Spranger, T; Bobbink, R. (2010). Use of dynamic soil-vegetation
models to assess impacts of nitrogen deposition on plant species composition: An
overview. Ecol Appl 20: 60-79. http://dx.doi.Org/10.1890/08-1019.l.
Dean, SL; Farrer, EC: Taylor, D; Porras-Alfaro, A: Suding, KN: Sinsabaugh, RL. (2014).
Nitrogen deposition alters plant-fungal relationships: Linking belowground dynamics to
aboveground vegetation change. Mol Ecol 23: 1364-1378.
http://dx.doi.Org/10.l 111/mec. 12541.
DeBusk, TA: Dierberg, FE: DeBusk, WF: Jackson, SD: Potts, JA: Galloway, SC: Sierer-Finn, D:
Gu, B. (2015). Sulfide concentration effects on Typha domingensis Pers. (cattail) and
Cladium jamaicense Crantz (sawgrass) growth in Everglades marshes. Aquat Bot 124:
78-84. http://dx.doi.Org/10.1016/i.aquabot.2015.03.006.
Declevre, H: Heylen, K, im: Van Colen, C: Willems, A. (2015). Dissimilatory nitrogen reduction
in intertidal sediments of a temperate estuary: small scale heterogeneity and novel nitrate-
to-ammonium reducers. FMICB 6: 1124. http://dx.doi.org/10.3389/fmicb.2015.01124.
Deegan, LA; Johnson, DS; Warren, RS; Peterson, BJ; Fleeger, JW: Fagherazzi, S; Wollheim,
WM. (2012). Coastal eutrophication as a driver of salt marsh loss. Nature 490: 388-+.
http://dx.doi.org/10.1038/naturel 1533.
DeHaves, DH: Schaberg, PG: Hawlev, GJ: Strimbeck, GR. (1999). Acid rain impacts on calcium
nutrition and forest health. Bioscience 49: 789-800. http://dx.doi.org/10.2307/1313570.
Deininger, A: Faithfull. CL: Bergstrom, AK. (2017a). Phytoplankton response to whole lake
inorganic N fertilization along a gradient in dissolved organic carbon. Ecology 98: 982-
994.
June 2018
R-52
DRAFT: Do Not Cite or Quote

-------
Deininger, A; Faithful!. CL; Karlsson. J; Klaus. M; Bergstrom, AK. (2017b). Pelagic food web
response to whole lake N fertilization. Limnol Oceanogr 62: 1498-1511.
http://dx.doi.org/10.10Q2/lno.10513.
Delgado-Baquerizo. M; Maestre. FT: Reich. PB; Jeffries. TC: Gaitan. JJ; Encinar. D; Berdugo.
M; Campbell. CD: Singh. BK. (2016). Microbial diversity drives multifunctionality in
terrestrial ecosystems. Nat Commun 7: 10541. http://dx.doi.org/10.1038/ncommslQ541.
Demalach. N: Zaadv. E: Kadmon. R. (2017). Light asymmetry explains the effect of nutrient
enrichment on grassland diversity [Letter], Ecol Lett 20: 60-69.
http://dx.doi.org/10. Ill 1/ele. 12706.
Dennis. RL; Mathur. R; Pleim. JE; Walker. JT. (2010). Fate of ammonia emissions at the local to
regional scale as simulated by the Community Multiscale Air Quality model. Atmos
Pollut Res 1: 207-214. http://dx.doi.org/10.5094/APR.2010.Q27.
Dennis, RL; Schwede, DB; Bash, JO: Pleim, JE; Walker, JT; Foley, KM. (2013). Sensitivity of
continental United States atmospheric budgets of oxidized and reduced nitrogen to dry
deposition parametrizations. Philos Trans R Soc Lond B Biol Sci 368: 20130124
(20130121-20130110). http://dx.doi.org/10.1098/rstb.2013.0124.
Desalme, D; Binet, P; Bernard, N; Gilbert, D; Toussaint, ML; Chiapusio, G. (201 la).
Atmospheric phenanthrene transfer and effects on two grassland species and their root
symbionts: A microcosm study. Environ Exp Bot 71: 146-151.
http://dx.doi.Org/10.1016/i.envexpbot.2010.l 1.009.
Desalme, D; Binet, P; Chiapusio, G. (2013). Challenges in tracing the fate and effects of
atmospheric polycyclic aromatic hydrocarbon deposition in vascular plants [Review],
Environ Sci Technol 47: 3967-3981. http://dx.doi.org/10.1021/es304964b.
Desalme, D; Binet, P; Epron, D; Bernard, N: Gilbert, D; Toussaint, ML: Plain, C: Chiapusio, G.
(201 lb). Atmospheric phenanthrene pollution modulates carbon allocation in red clover
(Trifolium pratense L.). Environ Pollut 159: 2759-2765.
http://dx.doi.Org/10.1016/i.envpol.2011.05.015.
Desalme, D; Chiapusio, G: Bernard, N: Gilbert, D; Toussaint, ML: Binet, P. (2012). Arbuscular
Mycorrhizal Fungal Infectivity in Two Soils as Affected by Atmospheric Phenanthrene
Pollution. Water Air Soil Pollut 223: 3295-3305. http://dx.doi.org/10.1007/sll270-Q12-
1110-z.
Desrosiers, M; Planas, D; Mucci, A. (2006). Mercury methylation in the epilithon of boreal
shield aquatic ecosystems. Environ Sci Technol 40: 1540-1546.
http://dx.doi.org/10.1021/es0508828.
Dettinger, MP: Cavan, DR. (1995). Large-scale atmospheric forcing of recent trends toward
early snowmelt runoff in California. J Clim 8: 606-623. http://dx.doi.Org/10.l 175/1520-
0442(1995)008<0606:LSAFOR>2.0.CQ;2.
Dev, S: Patra, AK: Mukheriee, A: Bhattacharya, J. (2015). Suitability of different growth
substrates as source of nitrogen for sulfate reducing bacteria. Biodegradation 26: 415-
430. http://dx.doi.org/10.1007/slQ532-015-9745-2.
June 2018
R-53
DRAFT: Do Not Cite or Quote

-------
Devaraiu. N; Govindasamy. B; Caldeira. K; Nemani. R. (2016). A model based investigation of
the relative importance of C02-fertilization, climate warming, nitrogen deposition and
land use change on the global terrestrial carbon uptake in the historical period. Clim
Dynam 47: 173-190. http://dx.doi.org/10.1007/sQ0382-015-2830-8.
Pevito. KJ: Dillon. PJ; Lazerte. BP. (1989). Phosphorus and nitrogen retention in five
Precambrian shield wetlands. Biogeochemistry 8: 185-204.
http://dx.doi.org/10.1007/BF000Q2888.
Devlin. M; Bricker. S: Painting. S. (2011). Comparison of five methods for assessing impacts of
nutrient enrichment using estuarine case studies. Biogeochemistry 106: 177-205.
http://dx.doi.org/10.1007/slQ533-011-9588-9.
Dewalle. PR: Bover. EW: Buda. AR. (2016). Exploring lag times between monthly atmospheric
deposition and stream chemistry in Appalachian forests using cross-correlation. Atmos
Environ 146: 206-214. http://dx.doi.Org/10.1016/i.atmosenv.2016.09.015.
Peyton. EB; Schwartz. JS: Robinson. RB; Neff. KJ: Moore. SE; Kulp. MA. (2009).
Characterizing episodic stream acidity during stormflows in the Great Smoky Mountains
National Park. Water Air Soil Pollut 196: 3-18. http://dx.doi.org/10.1007/sl 1270-008-
9753-5.
Pezi. S: Medlyn. BE: Tonon. G: Magnani. F. (2010). The effect of nitrogen deposition on forest
carbon sequestration: A model-based analysis. Global Change Biol 16: 1470-1486.
http://dx.doi.org/10.111 l/i.l365-2486.2009.02102.x.
Pias, T; Clemente, A; Martins-Loucao, MA; Sheppard, L; Bobbink, R; Cruz, C. (2014).
Ammonium as a driving force of plant diversity and ecosystem functioning: observations
based on 5 years' manipulation of N dose and form in a Mediterranean ecosystem. PLoS
ONE 9: e92517. http://dx.doi.org/10.1371/iournal.pone.0Q92517.
Piaz, G; Carrillo, C; Honrubia, M. (2010). Mycorrhization, growth and nutrition of Pinus
halepensis seedlings fertilized with different doses and sources of nitrogen. Annals of
Forest Science 67: 405-405. http://dx.doi.org/10.1051/forest/2009125.
Piaz. RJ; Eriksson-Hagg. H: Rosenberg. R. (2013). Hypoxia. In KJ Noone; UR Sumaila; RJ
Piaz (Eds.), Managing ocean environments in a changing climate : Sustainability and
economic perspectives (1st ed., pp. 67-96). Burlington, MA: Elsevier.
http://dx.doi.org/10.1016/B978-0-12-407668-6.000Q4-5.
Piaz. RJ: Rosenberg. R. (2008). Spreading dead zones and consequences for marine ecosystems
[Review], Science 321: 926-929. http://dx.doi.Org/10.l 126/science.l 156401.
Pieleman. WIJ; Vicca. S: Piikstra. FA: Hagedorn. F; Hovenden. MJ; Larsen. KS: Morgan. JA;
Voider. A: Beier. C: Pukes. JS: King. J: Leuzinger. S: Linder. S: Luo. Y; Oren. R; de
Angelis. P; Tingev. P; Hoosbeek. MR: Janssens. IA. (2012). Simple additive effects are
rare: A quantitative review of plant biomass and soil process responses to combined
manipulations of C02 and temperature. Global Change Biol 18: 2681-2693.
http://dx.doi.org/10.111 l/i,1365-2486.2012.02745.x.
Pierberg. FE; Brezonik. PL. (1983). Nitrogen and phosphorus mass balances in natural and
sewage-enriched cypress domes. J Appl Ecol 20: 323-337.
http://dx.doi.org/10.2307/2403395.
June 2018
R-54
DRAFT: Do Not Cite or Quote

-------
Dietze. MC; Moorcroft PR. (2011). Tree mortality in the eastern and central United States:
Patterns and drivers. Global Change Biol 17: 3312-3326.
http://dx.doi.org/10.1111/j. 1365-2486.2011,02477.x.
Dijkstra. FA. (2009). Modeling the flow of N-15 after a N-15 pulse to study long-term N
dynamics in a semiarid grassland. Ecology 90: 2171-2182. http://dx.doi.org/10.1890/08-
1172.1.
Dijkstra. FA: Adams. MA. (2015). Fire Eases Imbalances of Nitrogen and Phosphorus in Woody
Plants. Ecosystems 18: 769-779. http://dx.doi.org/10.1007/slQ021-015-9861-l.
Dillon, PJ; Reid, RA; Girard, R. (1986). Changes in the chemistry of lakes near Sudbury, Ontario
following reductions of S02 emissions. Water Air Soil Pollut 31: 59-65.
http://dx.doi.org/10.1007/BF0063Q819.
Dirnbock. T; Grandin. U; Bernhardt-Roemermann. M; Beudert. B; Canullo. R; Forsius. M;
Grabner, M, -T; Holmberg, M; Kleemola, S; Lundin, L; Mirtl, M; Neumann, M; Pompei,
E; Salemaa. M; Starlinger. F; Staszewski. T; Uzieblo. AK. (2014). Forest floor vegetation
response to nitrogen deposition in Europe. Global Change Biol 20: 429-440.
http://dx.doi.org/10. Ill 1/gcb. 12440.
Pise, NB; Ashmore, M; Belyazid, S; Bleeker, A; Bobbink, R; de Vries, W; Erisman, JW;
Spranger, T; van den Berg, L; Stevens, CJ. (2011). Nitrogen as a threat to European
terrestrial biodiversity. In MA Sutton; CM Howard; JW Erisman; G Billen; A Bleeker; P
Grennfelt; H van Grinsven; B Grizzetti (Eds.), The European Nitrogen Assessment (pp.
463-494). Cambridge, UK: Cambridge University Press.
http://dx.doi.org/10.1017/CBQ9780511976988.
Pise, NB: Rothwell, JJ; Gauci, V: van Per Salm, C: de Vries, W. (2009). Predicting dissolved
inorganic nitrogen leaching in European forests using two independent databases. Sci
Total Environ 407: 1798-1808. http://dx.doi.Org/10.1016/i.scitotenv.2008.l 1.003.
Pittman, JA; Shanlev, JB; Priscoll, CT; Aiken, GR; Chalmers, A, nnT; Towse, JE; Selvendiran,
P. (2010). Mercury dynamics in relation to dissolved organic carbon concentration and
quality during high flow events in three northeastern US streams. Water Resour Res 46.
http://dx.doi.org/10.1029/2009WR0Q8351.
Pivers, MT; Elliott, EM: Bain, PJ. (2014). Quantification of nitrate sources to an urban stream
using dual nitrate isotopes. Environ Sci Technol 48: 10580-10587.
http://dx.doi.org/10.1021/es404880i.
Pixon, RL; Grecay, PA; Targett, TE. (2017). Responses of juvenile Atlantic silverside, striped
killifish, mummichog, and striped bass to acute hypoxia and acidification: Aquatic
surface respiration and survival. Exp Mar Bio Ecol 493: 20-30.
http://dx.doi.Org/10.1016/i.iembe.2017.04.001.
Poak, PF; Bigger, P; Harding, EK; Marvier, MA; O'Malley, RE; Thomson, P. (1998). The
statistical inevitability of stability-diversity relationships in community ecology. Am Nat
151: 264-276. http://dx.doi.org/10.1086/286117.
June 2018
R-55
DRAFT: Do Not Cite or Quote

-------
Dodds, WK; Bouska, W; Eitzmann, JL; Pilger, TJ; Pitts. KL; Riley. AJ; Schloesser, JT;
Thornbrugh, DJ. (2009). Eutrophication of US Freshwaters: Analysis of Potential
Economic Damages. Environ Sci Technol 43: 12-19.
http://dx.doi.org/10.1021/es801217q.
Dodds. WK: Smith. VH. (2016). Nitrogen, phosphorus, and eutrophication in streams. Inland
Waters 6: 155-164. http://dx.doi.Org/10.5268/IW-6.2.909.
Dolislager. LJ; VanCuren. R; Pederson. J. . R.; Lashgari. A: McCaulev. E. (2012). A summary of
the Lake Tahoe Atmospheric Deposition Study (LTADS). Atmos Environ 46: 618-630.
http://dx.doi.Org/10.1016/i.atmosenv.2009.09.020.
Pollard. GJ: Unsworth. MH; Harve. MJ. (1983). Pollutant transfer in upland regions by occult
precipitation. Nature 302: 241-243. http://dx.doi.org/10.1038/302241a0.
Donagi. AE; Goren. AI. (1979). Use of indicator plants to evaluate atmospheric levels of
nitrogen dioxide in the vicinity of a chemical plant. Environ Sci Technol 13: 986-989.
http://dx.doi.org/10.1021/es60156a0Q3.
Donald. DB; Bogard. MJ: Finlav. K; Bunting. L; Leavitt. PR. (2013). Phytoplankton-specific
response to enrichment of phosphorus-rich surface waters with ammonium, nitrate, and
urea. PLoS ONE 8: e53277. http://dx.doi.org/10.1371/iournal.pone.0053277.
Donald. DB: Bogard. MJ: Finlav. K; Leavitt. PR. (2011). Comparative effects of urea,
ammonium, and nitrate on phytoplankton abundance, community composition, and
toxicity in hypereutrophic freshwaters. Limnol Oceanogr 56: 2161-2175.
http://dx.doi.Org/10.4319/lo.2011.56.6.2161.
Donev. SC. (2010). The growing human footprint on coastal and open-ocean biogeochemistry
[Review], Science 328: 1512-1516. http://dx.doi.Org/10.l 126/science.l 185198.
Donev. SC: Fabry. VJ; Feelv. RA; Klevpas. JA. (2009). Ocean acidification: The other C02
problem. Ann Rev Mar Sci 1: 169-192.
http://dx.doi.Org/10.l 146/annurev.marine.010908.163834.
Donev, SC: Ruckelshaus, M; Duffy, JE; Barry, JP; Chan, F; English, CA; Galindo, HM;
Grebmeier, JM; Hollowed, AB; Knowlton, N: Polovina, J: Rabalais, NN; Svdeman, WJ:
Talley, LP. (2012). Climate change impacts on marine ecosystems [Review], Ann Rev
Mar Sci 4: 11-37. http://dx.doi.Org/10.l 146/annurev-marine-041911-111611.
Dortch, Q: Whitledge, TE. (1992). Does nitrogen or silicon limit phytoplankton production in the
Mississippi River plume and nearby regions? Cont Shelf Res 12: 1293-1309.
http://dx.doi.org/10.1016/0278-4343(92N)90065-R.
Douglass, JG: France, KE; Richardson, JP: Duffy, JE. (2010). Seasonal and interannual change
in a Chesapeake Bay eelgrass community: Insights into biotic and abiotic control of
community structure. Limnol Oceanogr 55: 1499-1520.
http://dx.doi.Org/10.4319/lo.2010.55.4.1499.
Drake, JE: Stov, PC: Jackson, RB; Delucia, EH. (2008). Fine-root respiration in a loblolly pine
(Pinus taeda L.) forest exposed to elevated CO(2) and N fertilization. Plant Cell Environ
31: 1663-1672. http://dx.doi.Org/10.llll/i.1365-3040.2008.01869.x.
June 2018
R-56
DRAFT: Do Not Cite or Quote

-------
Drenner. RW; Chumchal, MM; Wente. SP; McGuire. M; Drenner. SM. (2011). Landscape-level
patterns of mercury contamination of fish in North Texas, USA. Environ Toxicol Chem
30: 2041-2045. http://dx.doi.org/10.10Q2/etc.589.
Drevnick. PE; Canfield. DE: Gorski. PR: Shinneman. ALC; Engstrom, PR: Muir. DCG: Smith.
GR; Garrison. PJ; Cleckner. LB: Hurley. JP: Noble. RB: Otter. RR: Oris. JT. (2007).
Deposition and cycling of sulfur controls mercury accumulation in isle Royale fish.
Environ Sci Technol 41: 7266-7272. http://dx.doi.org/10.1021/esQ712322.
Driscoll. C: Whitall. D; Aber. J: Bover. E; Castro. M; Cronan. C: Goodale. C: Groffman. P;
Hopkinson. C: Lambert. K; Lawrence. G: Ollinger. S. (2003a). Nitrogen pollution:
Sources and consequences in the US Northeast. Environment 45: 8-22.
Driscoll. CT; Cirmo. CP: Fahev. TJ; Blette. VL; Bukaveckas. PA: Burns. DA: Gubala. CP:
Leopold. DJ: Newton. RM; Ravnald. DJ: Schofield. CL; Yavitt. JB; Porcella. DB. (1996).
The Experimental Watershed Liming Study: Comparison of lake and watershed
neutralization strategies. Biogeochemistry DORDRECHT: 143-174.
Driscoll. CT; Driscoll. KM; Fakhraei. H; Civerolo. K. (2016). Long-term temporal trends and
spatial patterns in the acid-base chemistry of lakes in the Adirondack region of New York
in response to decreases in acidic deposition. Atmos Environ 146: 5-14.
http://dx.doi.Org/10.1016/i.atmosenv.2016.08.034.
Driscoll. CT; Driscoll. KM; Mitchell. MJ; Ravnal. DJ. (2003b). Effects of acidic deposition on
forest and aquatic ecosystems in New York State. Environ Pollut 123: 327-336.
Driscoll. CT; Driscoll. KM; Roy. KM; Dukett. J. (2007a). Changes in the chemistry of lakes in
the Adirondack Region of New York following declines in acidic deposition. Appl
Geochem 22: 1181-1188. http://dx.doi.Org/10.1016/i.apgeochem.2007.03.009.
Driscoll, CT; Driscoll, KM; Roy, KM; Mitchell, MJ. (2003c). Chemical response of lakes in the
Adirondack region of New York to declines in acidic deposition. Environ Sci Technol 37:
2036-2042. http://dx.doi.org/10.1021/es020924h.
Driscoll, CT; Fallon-Lambert, K; Chen, L. (2007b). Acidic deposition: Sources and ecological
effects. In GR Visgilio; DM Whitelaw (Eds.), Acid in the Environment: Lessons Learned
and Future Prospects, Part I (pp. 27-58). New York, NY: Springer.
http://dx.doi.org/10.1007/978-0-387-37562-5 3.
Driscoll, CT; Johnson, NM; Likens, GE; Feller, MC. (1988). Effects of acidic deposition on the
chemistry of headwater streams: a comparison between Hubbard Brook, New Hampshire,
and Jamieson Creek, British Columbia. Water Resour Res 24: 195-200.
http ://dx. doi .org/10.1029/WR024i002p00195.
Driscoll, CT; Lawrence, GB; Bulger, AJ; Butler, TJ; Cronan, CS; Eagar, C; Lambert, KF;
Likens, GE; Stoddard, JL; Weather, KC. (2001a). Acid rain revisited: Advances in
scientific understanding since the passage of the 1970 and 1990 Clean Air Act
Amendments. Hanover, NH: Hubbard Brook Research Foundation.
June 2018
R-57
DRAFT: Do Not Cite or Quote

-------
Driscoll. CT; Lawrence. GB; Bulger. AJ; Butler. TJ; Cronan. CS; Eagar, C; Lambert. KF;
Likens. GE; Stoddard. JL; Weathers. KC. (2001b). Acidic deposition in the northeastern
United States: Sources and inputs, ecosystem effects, and management strategies.
Bioscience 51: 180-198. http://dx.doi.org/10.1641/00Q6-
3568(200 n05ir0180:ADITNU12.0.CQ:2.
Driscoll. CT: Lehtinen. MP: Sullivan. TJ. (1994). Modeling the acid-base chemistry of organic
solutes in Adirondack, New York, lakes. Water Resour Res 30: 297-306.
Driscoll. CT: Newton. RM. (1985). Chemical characteristics of Adirondack lakes. Environ Sci
Technol 19: 1018-1024. http://dx.doi.org/10.1021/esQ0141a604.
Driscoll. CT: Postek. KM: Kretser. W: Ravnal. DJ. (1995). Long-term trends in the chemistry of
precipitation and lake water in the Adirondack region of New York, USA. Water Air Soil
Pollut 85: 583-588. http://dx.doi.org/10.1007/BF0Q476892.
Driscoll, CT; Whitall, D; Aber, J: Boyer, E; Castro, M; Cronan, C; Goodale, CL; Groffman, P;
Hopkinson, C: lambert, K: Lawrence, G: Ollinger, S. (2003d). Nitrogen pollution in the
northeastern United States: Sources, effects, and management options. Bioscience 53:
357-374.
Drott, A; Lambertsson, L; Bjorn, E; Skyllberg, U. (2008). Do potential methylation rates reflect
accumulated methyl mercury in contaminated sediments? Environ Sci Technol 42: 153-
158. http://dx.doi.org/10.1021/es0715851.
Drummond, JD; Bernal, S: von Schiller, D; Marti, E. (2016). Linking in-stream nutrient uptake
to hydrologic retention in two headwater streams. Freshw Sci 35: 1176-1188.
http://dx.doi.org/10.1086/688599.
Du, E. (2017). Integrating species composition and leaf nitrogen content to indicate effects of
nitrogen deposition. Environ Pollut 221: 392-397.
http://dx.doi.Org/10.1016/i.envpol.2016.12.001.
Du, E; Fang, J. (2014). Weak growth response to nitrogen deposition in an old-growth boreal
forest. Ecosphere 5: 1-9. http://dx.doi.Org/10.1890/ES14-00109.l.
Du, E; Liu, X: Fang, J. (2014a). Effects of nitrogen additions on biomass, stoichiometry and
nutrient pools of moss Rhytidium rugosum in a boreal forest in Northeast China. Environ
Pollut 188: 166-171. http://dx.doi.Org/10.1016/i.envpol.2014.02.011.
Du, Z; Wang, W; Zeng, W; Zeng, H. (2014b). Nitrogen Deposition Enhances Carbon
Sequestration by Plantations in Northern China. PLoS ONE 9: e87975.
http://dx.doi.org/10.1371/iournal.pone.0087975.
Duarte, CM. (2013). The role of seagrass in climate change mitigation and adaptation.
Presentation presented at The University of Western Australia and Spanish National
Research Council, October 24 -25, 2013, Bonn, Germany.
Duarte, CM: Boria, A: Carstensen, J: Elliott, M; Krause-Jensen, D; Marba, N. (2015). Paradigms
in the recovery of estuarine and coastal ecosystems. Estuaries Coasts 38: 1202-1212.
http://dx.doi.org/10.1007/sl2237-013-975Q-9.
June 2018
R-58
DRAFT: Do Not Cite or Quote

-------
Duarte. CM; Con lev. DJ; Carstensen. J; Sanchez-Cam acho. M. (2009). Return to Neverland:
Shifting Baselines Affect Eutrophication Restoration Targets. Estuaries Coasts 32: 29-36.
http://dx.doi.org/10.1007/sl2237-0Q8-9111-2.
Duarte. N: Pardo. LH; Robin-Abbott. MJ. (2013). Susceptibility of forests in the northeastern
USA to nitrogen and sulfur deposition: Critical load exceedance and forest health. Water
Air Soil Pollut 224: 1355. http://dx.doi.org/10.1007/sll270-Q12-1355-6.
Dubrovsky. NM; Burow. KR; Clark. GM; Gronberg. JM; PA. H; Hitt. KJ; Mueller. DK; Munn.
MP: Nolan. BT; Puckett. LJ; Rupert. MG: Short. TM; Spahr. NE; Sprague. LA: Wilber.
WG. (2010). Nutrients in the nation's streams and groundwater, 1992-2004. (Circular
1350). Reston, VA: U.S. Geological Survey.
http://water.usgs.gov/nawqa/nutrients/pubs/circl350/.
Duchesne. L; Houle. D. (2008). Impact of nutrient removal through harvesting on the
sustainability of the boreal forest. Ecol Appl 18: 1642-1651. http://dx.doi.org/10.1890/07-
1035.1.
Duchesne. L; Ouimet. R. (2009). Present-day expansion of American beech in northeastern
hardwood forests: Does soil base status matter? Can J For Res 39: 2273-2282.
http://dx.doi.Org/10.l 139/X09-172.
Duffy. JE; Cardinale. BJ; France. KE; Mcintyre. PB; Thebault. E; Loreau. M. (2007). The
functional role of biodiversity in ecosystems: incorporating trophic complexity. Ecol Lett
10: 522-538.
Duffy, WG: Kahara, SN. (2011). Wetland ecosystem services in California's Central Valley and
implications for the Wetland Reserve Program. Ecol Appl 21: S18-S30.
http://dx.doi.Org/10.1890/09-1338.l.
Duguma. D; Walton. WE. (2014). Effects of nutrients on mosquitoes and an emergent
macrophyte, Schoenoplectus maritimus, for use in treatment wetlands. J Vector Ecol 39:
1-13. http://dx.doi.Org/10.llll/i.1948-7134.2014.12063.x.
Dumont. E; Harrison. JA; Kroeze. C: Bakker. EJ; Seitzinger. SP. (2005). Global distribution and
sources of dissolved inorganic nitrogen export to the coastal zone: Results from a
spatially explicit, global model. Global Biogeochem Cycles 19: GB4S02.
http://dx.doi.org/10.1029/2005GB0Q2488.
Duncan. JM: Groffman. PM; Band. LE. (2013). Towards closing the watershed nitrogen budget:
Spatial and temporal scaling of denitrification. 118: 1105-1119.
http://dx.doi.org/10.1002/igrg.2009Q.
Dunck. B; Lima-Fernandes. E; Cassio. F; Cunha. A: Rodrigues. L; Pascoal. C. (2015).
Responses of primary production, leaf litter decomposition and associated communities
to stream eutrophi cation. Environ Pollut 202: 32-40.
http://dx.doi.Org/10.1016/i.envpol.2015.03.014.
Dunnette. PV: Higuera. PE; Mclauchlan. KK; Derr. KM: Briles. CE; Keefe. MH. (2014).
Biogeochemical impacts of wildfires over four millennia in a Rocky Mountain subalpine
watershed. NewPhytol 203: 900-912. http://dx.doi.org/10. Ill 1/nph. 12828.
June 2018
R-59
DRAFT: Do Not Cite or Quote

-------
Dupont J: Clair. TA; Gagnon, C; Jeffries. DS; Kahl. JS; Nelson. SJ; Peckenham. JM. (2005).
Estimation of critical loads of acidity for lakes in northeastern United States and eastern
Canada. Environ Monit Assess 109: 275-291. http://dx.doi.org/10.1007/slQ661-005-
6286-x.
Dupre. C: Stevens. CJ: Ranke. T; Bleeker. A: Peppler-Lisbach. C: Gowing, DJG: Pise. NB;
Dorland. E. du: Bobbink. R; Diekmann. M. (2010). Changes in species richness and
composition in European acidic grasslands over the past 70 years: The contribution of
cumulative atmospheric nitrogen deposition. Global Change Biol 16: 344-357.
http://dx.doi.org/10.111 l/i.l365-2486.2009.01982.x.
Duran. J: Morse. JL; Groffman. PM; Campbell. JL; Christenson. LM; Driscoll. CT; Fahev. TJ;
Fisk, MC; Likens, GE; Melillo, JM; Mitchell, MJ; Templer, PH; Vadeboncoeur, MA.
(2016). Climate change decreases nitrogen pools and mineralization rates in northern
hardwood forests. Ecosphere 7: 1-13. http ://dx.doi.org/10.1002/ecs2.1251.
DW, J: SE, L. (1992). Atmospheric deposition and forest nutrient cycling: A synthesis of the
integrated forest study. In DW Johnson; SE Lindberg (Eds.). New York, NY: Springer-
Verlag, Inc.
Eadie, BJ; McKee, BA; Lansing, MB; Robbins, JA; Metz, S; Trefry, JH. (1994). Records of
nutrient-enhanced coastal ocean productivity in sediments from the Louisiana continental
shelf. Estuaries Coasts 17: 754-765. http://dx.doi.org/10.2307/1352745.
Eagles-Smith, CA; Herring, G; Johnson, B; Graw, R. (2016). Conifer density within lake
catchments predicts fish mercury concentrations in remote subalpine lakes. Environ
Pollut 212: 279-289. http://dx.doi.Org/10.1016/i.envpol.2016.01.049.
Eastaugh, CS; Poetzelsberger, E; Hasenauer, H. (2011). Assessing the impacts of climate change
and nitrogen deposition on Norway spruce (Picea abies L. Karst) growth in Austria with
BIOME-BGC. Tree Physiol 31: 262-274. http://dx.doi.org/10.1093/treephys/tpr033.
Ecklev, CS; Hintelmann, H. (2006). Determination of mercury methylation potentials in the
water column of lakes across Canada. Sci Total Environ 368: 111-125.
http://dx.doi.Org/10.1016/i.scitotenv.2005.09.042.
Ecklev, CS; Luxton, TP; Goetz, J; Mckernan, J. (2017). Water-level fluctuations influence
sediment porewater chemistry and methylmercury production in a flood-control reservoir.
Environ Pollut 222: 32-41. http://dx.doi.Org/10.1016/i.envpol.2017.01.010.
Ecklev, CS; Luxton, TP; Mckernan, JL; Goetz, J; Goulet, J. (2015). Influence of reservoir water
level fluctuations on sediment methylmercury concentrations downstream of the
historical Black Butte mercury mine, OR. Appl Geochem 61: 284-293.
http://dx.doi.Org/10.1016/i.apgeochem.2015.06.011.
Edmondson, WT. (1969). Eutrophication in North America. In Eutrophication: Causes,
Consequences, Correctives (Proceedings from a Symposium). Washington, DC: National
Academy of Sciences.
Edmondson, WT. (1991). The uses of ecology: Lake Washington and beyond. In The Uses of
Ecology: Lake Washington and Beyond. Seattle, WA: University of Washington Press.
June 2018
R-60
DRAFT: Do Not Cite or Quote

-------
Edokpa. DA; Evans. MG; Roth well. JJ. (2015). High fluvial export of dissolved organic nitrogen
from a peatland catchment with elevated inorganic nitrogen deposition. Sci Total Environ
532: 711-722. http://dx.doi.Org/10.1016/i.scitotenv.2015.06.072.
Edwards. IP: Zak. PR: Kellner. H; Eisenlord. SD: Pregitzer. KS. (2011). Simulated atmospheric
N deposition alters fungal community composition and suppresses ligninolytic gene
expression in a northern hardwood forest. PLoS ONE 6: e20421.
http://dx.doi. org/ 10.1371/i ournal. pone.0020421.
Edwards. PJ. (1998). Sulfur cycling, retention, and mobility in soils: A review. (Gen. Tech. Rep.
NE-250). Radnor, PA: Department of Agriculture, Forest Service, Northeastern Research
Station.
Edwards. PJ: Helvev. JD. (1991). Long-term ionic increases from a central Appalachian forested
watershed. J Environ Qual 20: 250-255.
Eeva, T; Ahola, M; Lehikoinen. E. (2009). Breeding performance of blue tits (Cyanistes
caeruleus) and great tits (Parus major) in a heavy metal polluted area. Environ Pollut 157:
3126-3131. http://dx.doi.Org/10.1016/i.envpol.2009.05.040.
Eeva. T; Lehikoinen. E. (2015). Long-term recovery of clutch size and egg shell quality of the
pied flycatcher (Ficedula hypoleuca) in a metal polluted area. Environ Pollut 201: 26-33.
http://dx.doi.Org/10.1016/i.envpol.2015.02.027.
Eeva. T; Lehikoinen. E. sa. (2010). Polluted environment and cold weather induce laying gaps in
great tit and pied flycatcher. Oecologia 162: 533-539. http://dx.doi.org/10.1007/sQ0442-
009-1468-9.
Eeva. T; Penttinen. R. (2009). Leg deformities of oribatid mites as an indicator of environmental
pollution. Sci Total Environ 407: 4771-4776.
http://dx.doi.Org/10.1016/i.scitotenv.2009.05.013.
Eeva. T; Rainio. K; Suominen. O. (2010). Effects of pollution on land snail abundance, size and
diversity as resources for pied flycatcher, Ficedula hypoleuca. Sci Total Environ 408:
4165-4169. http://dx.doi.Org/10.1016/i.scitotenv.2010.05.028.
Eeva. T; Rainio. M; Kanerva. M; Salminen. JP. (2012). Plasma carotenoid levels are not directly
related to heavy metal exposure or reproductive success in three insectivorous passerines.
Environ Toxicol Chem 31: 1363-1369. http://dx.doi.org/10.1002/etc.1821.
Egan, S, Holzman, B. (2003). The biogeography of the lace lichen (R. menziesii). Available
online at http://online.sfsu.edU/bholzman/courses/Fall%2003%20proiect/R.menziesii.htm
(accessed March 16, 2016).
Egeiton-Warburton. LM; Allen. EB. (2000). Shifts in arbuscular mycorrhizal communities along
an anthropogenic nitrogen deposition gradient. Ecol Appl 10: 484-496.
http://dx.doi.org/10.2307/26411Q8.
Egerton-Warburton. LM: Johnson. NC: Allen. EB. (2007). Mycorrhizal community dynamics
following nitrogen fertilization: A cross-site test in five grasslands. Ecol Monogr 77: 527-
544.
June 2018
R-61
DRAFT: Do Not Cite or Quote

-------
Eggleston. EM; Lee. DY; Owens. MS; Cornwell. JC; Crump. BC; Hewson. I an. (2015). Key
respiratory genes elucidate bacterial community respiration in a seasonally anoxic
estuary. Environ Microbiol 17: 2306-2318. http://dx.doi.Org/10.l 111/1462-2920.12690.
Ehle. DS; Baker. WL. (2003). Disturbance and stand dynamics in ponderosa pine forests in
Rocky Mountain National Park, USA. Ecol Monogr 73: 543-566.
Eichert. T; Kurtz. A; Steiner. U; Goldbach. HE. (2008). Size exclusion limits and lateral
heterogeneity of the stomatal foliar uptake pathway for aqueous solutes and water-
suspended nanoparticles. Physiol Plant 134: 151-160. http://dx.doi.org/10.Ill 1/i.1399-
3054.2008.01135.x.
Eickenscheidt. N; Brumme. R. (2012). NOx and N20 fluxes in a nitrogen-enriched European
spruce forest soil under experimental long-term reduction of nitrogen depositions. Atmos
Environ 60: 51-58. http://dx.doi.Org/10.1016/i.atmosenv.2012.06.018.
Eilers, J; Vache, K; Eilers, B; Sweets, R; Cornett, J. (2016). Assessing the Chemical and
Biological Resilience of Lakes in the Cascade Range to Acidic Deposition. Water Air
Soil Pollut 227. http://dx.doi.org/10.1007/s 11270-016-3135-1.
Eilers. JM; Brakke. DF; Landers. DH; Overton. WS. (1989). Chemistry of lakes in designated
wilderness areas in the western United States. Environ Monit Assess 12: 3-21.
http://dx.doi.org/10.1007/BF0Q396724.
Eimers. MC; Watmough. SA. (2016). Increasing nitrate concentrations in streams draining into
Lake Ontario. J Great Lakes Res 42: 356-363.
http://dx.doi.Org/10.1016/i.iglr.2016.01.002.
Eimil-Fraga. C; Fernandez-Saniurio. MJ; Rodriguez-Soalleiro. R; Alvarez-Rodriguez. E. (2016).
Aluminium toxicity risk for Pinus pinaster in acid soils (Galicia, NW Spain). Land
Degrad Dev 27: 1731-1739. http://dx.doi.org/10.10Q2/ldr.2539.
Eisenhauer. N; Cesarz. S; Roller. R; Worm. K; Reich. PB. (2012). Global change belowground:
Impacts of elevated C02, nitrogen, and summer drought on soil food webs and
biodiversity. Global Change Biol 18: 435-447. http://dx.doi.org/10.1111/i.1365-
2486.2011.02555.x.
Eisenhauer. N; Dobies. T; Cesarz. S; Hobbie. SE; Meyer. RJ; Worm. K; Reich. PB. (2013). Plant
diversity effects on soil food webs are stronger than those of elevated C02 and N
deposition in a long-term grassland experiment. Proc Natl Acad Sci USA 110: 6889-
6894. http://dx.doi.org/10.1073/pnas.1217382110.
Eisenlord. SD; Freedman. Z; Zak. PR; Xue. K; He. Z; Zhou. J. (2013). Microbial Mechanisms
Mediating Increased Soil C Storage under Elevated Atmospheric N Deposition. Appl
Environ Microbiol 79: 1191-1199. http://dx.doi.Org/10.l 128/AEM.03156-12.
Eisenlord. SD; Zak. DR. (2010). Simulated atmospheric nitrogen deposition alters actinobacterial
community composition in forest soils. Soil Sci Soc Am J 74: 1157-1166.
http://dx.doi.org/10.2136/sssai2009.024Q.
June 2018
R-62
DRAFT: Do Not Cite or Quote

-------
Ekstrom. JA; Suatoni. L; Coolev. SR; Pendleton. LH; Waldbusser. GG; Cinner. JE; Ritter. J;
Langdon. C; van Hooidonk. R; Gledhill. D; Wellman. K; Beck. MW; Brander. LM;
Rittschof, D, an; Doherty, C; Edwards, PET; Portela, R. (2015). Vulnerability and
adaptation of US shellfisheries to ocean acidification. Nat Clim Chang 5: 207-214.
http://dx.doi.org/10.1038/NCLIMATE2508.
Eldridge. PM; Morse. JW. (2008). Origins and temporal scales of hypoxia on the Louisiana
shelf: Importance of benthic and sub-pycnocline water metabolism. Mar Chem 108: 159-
171. http://dx.doi.Org/10.1016/i.marchem.2007.l 1.009.
Eldridge. PM; Roelke. PL. (2011). 9.11-Hypoxia in waters of the coastal zone: Causes, effects,
and modeling approaches. In E Wolanski; D McLusky (Eds.), Treatise on Estuarine and
Coastal Science: Estuarine and Coastal Ecosystem Modelling (pp. 193-215). Amsterdam,
The Netherlands: Elsevier Inc. http://dx.doi.org/10.1016/B978-Q-12-374711-2.00911-6.
Elias. PE; Burger. JA; Adams. MB. (2009). Acid deposition effects on forest composition and
growth on the Monongahela National Forest, West Virginia. For Ecol Manage 258: 2175-
2182. http://dx.doi.Org/10.1016/i.foreco.2009.05.004.
Eliason. SA; Allen. EB. (1997). Exotic grass competition in suppressing native shrubland re-
establishment. RestorEcol 5: 245-255. http://dx.doi.Org/10.1046/i. 1526-
100X. 1997.09729.x.
Eliasson. PE; Agren. GI. (2011). Feedback from soil inorganic nitrogen on soil organic matter
mineralisation and growth in a boreal forest ecosystem. Plant Soil 338: 193-203.
http://dx.doi.org/10.1007/slll04-010-Q478-z.
Ellenberg. H. (1985). Veranderungen der floa mitteleuropas unter dem einfluss von dungung und
immissionen. Schweiz Z Forstwes 136: 19-39.
Elling. W; Dittmar. C; Pfaffelmoser. K; Rotzer. T. (2009). Dendroecological assessment of the
complex causes of decline and recovery of the growth of silver fir (Abies alba Mill.) in
Southern Germany. For Ecol Manage 257: 1175-1187.
http://dx.doi.Org/10.1016/i.foreco.2008.10.014.
Elliott. KH; Elliott. JE. (2016). Origin of Sulfur in Diet Drives Spatial and Temporal Mercury
Trends in Seabird Eggs From Pacific Canada 1968-2015. Environ Sci Technol 50:
13380-13386. http://dx.doi.org/10.1021/acs.est.6b05458.
Elliott. KJ; Knoepp. JD; Vose. JM; Jackson. WA. (2013). Interacting effects of wildfire severity
and liming on nutrient cycling in a southern Appalachian wilderness area. Plant Soil 366:
165-183. http://dx.doi.org/10.1007/slllQ4-012-1416-z.
Elliott. KJ; Vose. JM; Knoepp. JD; Johnson. DW; Swank. WT; Jackson. W. (2008). Simulated
effects of sulfur deposition on nutrient cycling in Class I wilderness areas. J Environ Qual
37: 1419-1431. http://dx.doi.org/10.2134/iea2007.0358.
Ellis. BK; Craft. JA; Stanford. JA. (2015). Long-term atmospheric deposition of nitrogen,
phosphorus and sulfate in a large oligotrophic lake. Peer J 3: e841.
http://dx.doi.org/10.7717/peeri.841.
June 2018
R-63
DRAFT: Do Not Cite or Quote

-------
Ellis. RA; Jacob. DJ; Sulprizio. MP; Zhang. L; Holmes. CD; Schichtel. BA; Blett T; Porter. E;
Pardo. LH; Lynch. JA. (2013). Present and future nitrogen deposition to national parks in
the United States: Critical load exceedances. Atmos Chem Phys 13: 9083-9095.
http://dx.doi.org/10.5194/acp-13-9083-2Q13.
Elser. JJ; Andersen. T; Baron. JS; Bergstrom. A. -K; Jansson. M; Kyle. M; Nvdick. KR; Steger.
L; Hessen. D. (2009a). Shifts in Lake N:P stoichiometry and nutrient limitation driven by
atmospheric nitrogen deposition. Science 326: 835-837.
http://dx.doi.Org/10.l 126/science.l 176199.
Elser. JJ; Bracken. MES; Cleland. EE; Gruner. DS; Harpole. WS; Hillebrand. H; Ngai. JT;
Seabloom. EW; Shurin. JB; Smith. JE. (2007). Global analysis of nitrogen and
phosphorus limitation of primary producers in freshwater, marine, and terrestrial
ecosystems. Ecol Lett 10: 1135-1142. http://dx.doi.Org/10.l 111/j. 1461-
0248.2007.01113.x.
Elser. JJ; Havakawa. K; Urabe. J. (2001). Nutrient limitation reduces food quality for
zooplankton: Daphnia response to seston phosphorus enrichment. Ecology 82: 898-903.
http://dx.doi.org/10.2307/2680208.
Elser. JJ; Kyle. M; Steger. L; Nvdick. KR; Baron. JS. (2009b). Nutrient availability and
phytoplankton nutrient limitation across a gradient of atmospheric nitrogen deposition.
Ecology 90: 3062-3073. http://dx.doi.Org/10.1890/08-1742.l.
Elser. JJ; Marzolf. ER; Goldman. CR. (1990). Phosphorus and nitrogen limitation of
phytoplankton growth in the freshwaters of North America: A review and critique of
experimental enrichments [Review], Can J Fish Aquat Sci 47: 1468-1477.
http://dx.doi.Org/10.l 139/P90-165.
Elser. JJ; Peace. AL; Kyle. M; Woiewodzic. M; McCrackin. ML; Andersen. T; Hessen. DO.
(2010). Atmospheric nitrogen deposition is associated with elevated phosphorus
limitation of lake zooplankton. Ecol Lett 13: 1256-1261.
http://dx.doi.Org/10.llll/i.1461-0248.2010.01519.x.
Elvidge. CK; Brown. GE. (2014). Predation costs of impaired chemosensory risk assessment on
acid-impacted juvenile Atlantic salmon (Salmo salar). Can J Fish Aquat Sci 71: 756-762.
http://dx.doi.Org/10.l 139/cifas-2013-0633.
Elvir. JA; Wiersma. GB; Bethers. S; Kenlan. P. (2010). Effects of chronic ammonium sulfate
treatment on the forest at the Bear Brook Watershed in Maine. Environ Monit Assess
171: 129-147. http://dx.doi.org/10.1007/s 10661-010-1523-3.
Elvir. JA; Wiersma. GB; Day. ME; Greenwood. MS; Fernandez. IJ. (2006). Effects of enhanced
nitrogen deposition on foliar chemistry and physiological processes of forest trees at the
Bear Brook Watershed in Maine. For Ecol Manage 221: 207-214.
http://dx.doi.Org/10.1016/i.foreco.2005.09.022.
Elvir. JA; Wiersma. GB; White. AS; Fernandez. IJ. (2003). Effects of chronic ammonium sulfate
treatment on basal area increment in red spruce and sugar maple at the Bear Brook
Watershed in Maine. Can J For Res 33: 861-869. http://dx.doi.Org/10.l 139/X03-016.
June 2018
R-64
DRAFT: Do Not Cite or Quote

-------
El wood. JW. (1991). Southeast overview. In DF Charles; S Christie (Eds.), Acidic Deposition
and Aquatic Ecosystems: Regional Case Studies (pp. 291-295). New York, NY:
Springer-Verlag. http://dx.doi.ore/10.1007/978-l-4613-9038-l 12.
El wood. JW: Sale. MJ; Kaufmann. PR; Cada. GF. (1991). The southern Blue Ridge province. In
DF Charles (Ed.), Acidic Deposition and Aquatic Ecosystems: Regional Case Studies
(pp. 319-364). New York, NY: Springer-Verlag New York, Inc.
http://dx.doi.org/10.1007/978-l-4613-9038-l 14.
Emmett. BA; Beier. C: Estiarte. M; Tietema. A: Kristensen. HL; Williams. D; Penuelas. J:
Schmidt. I: Sowerbv. A. (2004). The response of soil processes to climate change:
Results from manipulation studies of shrublands across an environmental gradient.
Ecosystems 7: 625-637. http://dx.doi.org/10.1007/sl0021-004-022Q-x.
Emmett. BA: Boxman. D; Bredemeier. M; Gunderson. P; Kionaas. OJ: Moldan. F; Schleppi. P;
Tietema. A: Wright. RF. (1998). Predicting the effects of atmospheric nitrogen deposition
in conifer stands: evidence from the NITREX ecosystem-scale experiments. Ecosystems
1: 352-360.
Enders. SK; Pagani. M; Pantoia. S: Baron. JS: Wolfe. AP; Pedentchouk. N: Nunez. L. (2008).
Compound-specific stable isotopes of organic compounds from lake sediments track
recent environmental changes in an alpine ecosystem, Rocky Mountain National Park,
Colorado. Limnol Oceanogr 53: 1468-1478.
Engardt. M; Langner. J. (2013). Simulations of future sulphur and nitrogen deposition over
Europe using meteorological data from three regional climate projections. Tellus B Chem
Phys Meteorol 65. http://dx.doi.org/10.3402/tellusb.v65i0.20348.
Enowashu. E; Poll. C: Lamersdorf. N: Kandeler. E. (2009). Microbial biomass and enzyme
activities under reduced nitrogen deposition in a spruce forest soil. Appl Soil Ecol 43: 11-
12. http://dx.doi.Org/10.1016/i.apsoil.2009.05.003.
Epstein. DM: Wurtsbaugh. WA; Baker. MA. (2012). Nitrogen partitioning and transport through
a subalpine lake measured with an isotope tracer. Limnol Oceanogr 57: 1503-1516.
http://dx.doi.Org/10.4319/lo.2012.57.5.1503.
Erisman, JW: Galloway, JN; Seitzinger, S; Bleeker, A; Pise, NB; Petrescu, AMR; Leach, AM;
de Vries, W. (2013). Consequences of human modification of the global nitrogen cycle.
Philos Trans R Soc Lond B Biol Sci 368: 1-9. http://dx.doi.org/10.1098/rstb.2Q13.0116.
Escobedo, FJ; Kroeger, T; Wagner, JE. (2011). Urban forests and pollution mitigation:
Analyzing ecosystem services and disservices [Review], Environ Pollut 159: 2078-2087.
http://dx.doi.Org/10.1016/i.envpol.2011.01.010.
Eshleman, KN: Kline, KM: Morgan, RP, II: Castro, NM; Neglev, TL. (2008). Contemporary
trends in the acid-base status of two acid-sensitive streams in western Maryland. Environ
Sci Technol 42: 56-61. http://dx.doi.org/10.1021/esQ71195e.
Eshleman, KN: Sabo, RD. (2016). Declining nitrate-N yields in the Upper Potomac River Basin:
What is really driving progress under the Chesapeake Bay restoration? Atmos Environ
146: 280-289. http://dx.doi.Org/10.1016/i.atmosenv.2016.07.004.
June 2018
R-65
DRAFT: Do Not Cite or Quote

-------
Eshleman. KN; Sabo. RD: Kline. KM. (2013). Surface water quality is improving due to
declining atmospheric N deposition. Environ Sci Technol 47: 12193-12200.
http://dx.doi.ore/10.1021/es4028748.
Esparza. ML: Farrell. AE; Craig. DJ: Swanson. C: Dhaliwal. BS: Berg. GM. (2014). Impact of
atypical ammonium concentrations on phytoplankton abundance and composition in
fresh versus estuarine waters. Aquatic Biology 21: 191-204.
http://dx.doi.org/10.3354/ab00588.
Esque. TC: Kave. JP; Eckert. SE; DeFalco. LA: Tracy. CR. (2010). Short-term soil inorganic N
pulse after experimental fire alters invasive and native annual plant production in a
Mojave Desert shrubland. Oecologia 164: 253-263. http://dx.doi.org/10.1007/sQ0442-
010-1617-1.
Esseen. PA: Ekstrom. M; Westerlund. B; Palmqvist. K; Jonsson. BG: Grafstrom. A: Stahl. G.
(2016). Broad-scale distribution of epiphytic hair lichens correlates more with climate
and nitrogen deposition than with forest structure. Can J For Res 40: 1348-1358.
http://dx.doi.org/10.1139/cifr-2016-0113.
Estop-Aragones. C: Zaiac. K; Blodau. C. (2016). Effects of extreme experimental drought and
rewetting on C02 and CH4 exchange in mesocosms of 14 European peatlands with
different nitrogen and sulfur deposition. Global Change Biol 22: 2285-2300.
http://dx.doi.org/10. Ill 1/gcb. 13228.
Etzold. S: Waldner. P; Thimonier. A: Schmitt. M; Dobbertin. M. (2014). Tree growth in Swiss
forests between 1995 and 2010 in relation to climate and stand conditions: Recent
disturbances matter. For Ecol Manage 311: 41-55.
http://dx.doi.Org/10.1016/i.foreco.2013.05.040.
Evans. CD: Chapman. PJ; Clark. JM; Monteith. DT. (2006). Alternative explanations for rising
dissolved organic carbon export form organic soils. Global Change Biol 12: 2044-2053.
http://dx.doi.Org/10.llll/i.1365-2486.2006.01241.x.
Evans. CD: Goodale. CL; Caporn. SJM; Pise. NB; Emmett. BA; Fernandez. IJ; Field. CD:
Findlav. SEG: Lovett. GM: Meesenburg. H: Moldan. F; Sheppard. LJ. (2008). Does
elevated nitrogen deposition or ecosystem recovery from acidification drive increased
dissolved organic carbon loss from upland soil? A review of evidence from field nitrogen
addition experiments. Biogeochemistry 91: 13-35. http://dx.doi.org/10.1007/slQ533-008-
9256-x.
Evans. CD: Malcolm. IA: Shilland. EM: Rose. NL: Turner. SD: Crilly. A: Norris. D: Granath. G:
Monteith. DT. (2017). Sustained Biogeochemical Impacts of Wildfire in a Mountain
Lake Catchment. Ecosystems 20: 813-829. http://dx.doi.org/10.1007/slQ021-016-0064-l.
Evans. CD: Monteith. DT. (2001). Chemical trends at lakes and streams in the UK Acid Waters
Monitoring Network, 1988-2000: Evidence for recent recovery at a national scale. Hydrol
Earth Syst Sci 5: 351-366. http://dx.doi.org/10.5194/hess-5-351-2Q01.
Evans. CD: Monteith. DT: Cooper. DM. (2005). Long-term increases in surface water dissolved
organic carbon: Observations, possible causes and environmental impacts. Environ Pollut
137: 55-71. http://dx.doi.Org/10.1016/i.envpol.2004.12.031.
June 2018
R-66
DRAFT: Do Not Cite or Quote

-------
Evans. JR. (1989). Photosynthesis and nitrogen relationships in leaves of C-3 plants. Oecologia
78: 9-19. http://dx.doi.org/10.1007/BF0Q377192.
Evans. MA: Sea via. D. (2013). Exploring estuarine eutrophi cation sensitivity to nutrient loading.
Limnol Oceanogr 58: 569-578. http://dx.doi.Org/10.4319/lo.2013.58.2.0569.
Everitt, JH; Drawe, PL; Lonard, RI. (1999). Field guide to the broad leaved herbaceous plants of
South Texas: Used by livestock and wildlife. In Field Guide to the Broad Leaved
Herbaceous Plants of South Texas: Used by Livestock and Wildlife. Lubbock, TX: Texas
Tech University Press.
Exley, C; Phillips, MJ. (1988). Acid rain: Implications for the farming of salmonids. In JF Muir;
RJ Roberts (Eds.), Recent Advances in Aquaculture (pp. 225-341). Dordrecht,
Netherlands: Springer Netherlands, http://dx.doi.org/10.1007/978-94-011-9743-4 4.
Facelli. JM; Pickett. STA. (1991). Plant litter: Its dynamics and effects on plant community
structure. Bot Rev 57: 1-32. http://dx.doi.org/10.1007/BFQ2858763.
Fakhraei. H; Driscoll. CT. (2015). Proton and Aluminum Binding Properties of Organic Acids in
Surface Waters of the Northeastern US. Environ Sci Technol 49: 2939-2947.
http://dx.doi.org/10.1021/es504024u.
Fakhraei. H; Driscoll. CT: Renfro. JR; Kulp. MA: Blett. TF; Brewer. PF; Schwartz. JS. (2016).
Critical loads and exceedances for nitrogen and sulfur atmospheric deposition in Great
Smoky Mountains National Park, United States. Ecosphere 7: 1-28.
http://dx.doi.org/10.1002/ecs2.1466.
Fakhraei. H; Driscoll. CT: Selvendiran. P; DePinto. JV: Bloomfield. J: Quinn. S: Rowell. HC.
(2014). Development of a total maximum daily load (TMDL) for acid-impaired lakes in
the Adirondack region of New York. Atmos Environ 95: 277-287.
http://dx.doi.Org/10.1016/i.atmosenv.2014.06.039.
Fakhraei. HK; Driscoll. CT: Kulp. MA: Renfro. JR: Blett. TF: Brewer. PF: Schwartz. JS.
(2017a). Sensitivity and uncertainty analysis of PnET-BGC to inform the development of
Total Maximum Daily Loads (TMDLs) of acidity in the Great Smoky Mountains
National Park. Environ Modell Softw 95: 156-167.
http://dx.doi.Org/10.1016/i.envsoft.2017.06.013.
Fakhraei. HK: Driscoll. CT: Kulp. MA: Renfro. JR: Blett. TF: Brewer. PF: Schwartz. JS.
(2017b). Sensitivity and uncertainty analysis of PnET-BGC to inform the development of
Total Maximum Daily Loads (TMDLs) of acidity in the Great Smoky Mountains
National Park: Supplementary materials [Supplemental Data], Environ Modell Softw 95.
Falkengren-Grerup. U. (1986). Soil acidification and vegetation changes in deciduous forest in
southern Sweden. Oecologia 70: 339-347. http://dx.doi.org/10.1007/BF0Q379494.
Falkengren-Grerup. U. (1989). Soil acidification and its impact on ground vegetation. Ambio 18:
179-183.
Falkengren-Grerup. U: Eriksson. H. (1990). Changes in soil, vegetation, and forest yield between
1947 and 1988 in beech and oak sites of southern Sweden. For Ecol Manage 38: 37-53.
http://dx.doi.org/10.1016/0378-l 127(90^)90084-0.
June 2018
R-67
DRAFT: Do Not Cite or Quote

-------
Fangmeier. A; Hadwiger-Fangmeier. A; Van der Eerden. L; Jager. H. -J. (1994). Effects of
atmospheric ammonia on vegetation—a review [Review], Environ Pollut 86: 43-82.
http ://dx. doi .org/10.1016/0269-7491 (94)90008-6.
Farahat EA; Linderholm. HW; Lechowicz. MJ. (2016). Influence of dust deposition and climate
on the radial growth of Tsuga canadensis near its northern range limit. European Journal
of Forest Research 135: 69-76. http://dx.doi.org/10.1007/slQ342-015-0917-8.
Farmer. AM: Bates. JW: Bell JNB. (1992). Ecophysiological effects of acid rain on bryophytes
and lichens. In JW Bates; AM Farmer (Eds.), Bryophytes and Lichens in a Changing
Environment. Oxford, UK: Claredon Press.
Farquhar. GD; Firth. PM; Wetselaar. R; Weir. B. (1980). On the gaseous exchange of ammonia
between leaves and the environment: Determination of the ammonia compensation point.
J Plant Physiol 66: 710-714. http://dx.doi.Org/10.l 104/pp.66.4.710.
Farrer, EC: Ashton, IW; Spasojevic, MJ: Fu, S; Gonzalez, DJX; Suding, KN. (2015). Indirect
effects of global change accumulate to alter plant diversity but not ecosystem function in
alpine tundra. J Ecol 103: 351-360. http://dx.doi.org/10.1111/1365-2745.12363.
Farrer. EC: Herman. DJ; Franzova. E; Pham. T; Suding. KN. (2013). Nitrogen deposition, plant
carbon allocation, and soil microbes: Changing interactions due to enrichment. Am J Bot
100: 1458-1470. http://dx.doi.org/10.3732/aib.1200513.
Farrer. EC: Suding. KN. (2016). Teasing apart plant community responses to N enrichment: The
roles of resource limitation, competition and soil microbes [Letter], Ecol Lett 19: 1287-
1296. http://dx.doi.org/10.Ill 1/ele. 12665.
Farrior. CE; Tilman. D; Dybzinski. R; Reich. PB; Levin. SA; Pacala. SW. (2013). Resource
limitation in a competitive context determines complex plant responses to experimental
resource additions. Ecology 94: 2505-2517. http://dx.doi.Org/10.1890/12-1548.l.
Fatemi. FR; Fernandez. IJ; Szillery. J: Norton. SA: Rustad. LE. (2012). Soil Solution Chemical
Response to Two Decades of Experimental Acidification at the Bear Brook Watershed in
Maine. Water Air Soil Pollut 223: 6171-6186. http://dx.doi.org/10.1007/sll270-Q12-
1348-5.
Feel v. RA; Sabine. CL; Hernandez-Avon. JM; Ianson. D; Hales. B. (2008). Evidence for
upwelling of corrosive "acidified" water onto the continental shelf. Science 320: 1490-
1492. http://dx.doi.org/10.1126/science. 1155676.
Felger. RS: Moser. MB. (1985). People of the desert and sea: Ethnobatany on the Seri Indians.
Tucson, AZ: University of Arizona Press.
Feller. IC: Lovelock. CE: Piou. C. (2009). Growth and nutrient conservation in rhizophora
mangle in response to fertilization along latitudinal and tidal gradients. Smithsonian
Contributions to Marine Science 38: 345-358.
Feng. MH: Shan. XQ: Zhang. SZ: Wen. B. (2005). Comparison of a rhizosphere-based method
with other one-step extraction methods for assessing the bioavailability of soil metals to
wheat. Chemosphere 59: 939-949. http://dx.doi.org/ 10.1016/j.chemosphere.2004.11.056.
June 2018
R-68
DRAFT: Do Not Cite or Quote

-------
Feng. S; Ai, Z; Zheng. S; Gu, B; Li. Y. (2014). Effects of dry out and inflow water quality on
mercury methylation in a constructed wetland. Water Air Soil Pollut 225: 1929.
http://dx.doi.org/10.1007/sll270-Q14-1929-6.
Fenn. ME: Allen. EB; Geiser. ALH. (201 la). Mediterranean California. In Assessment of
nitrogen deposition effects and empirical critical loads of nitrogen for ecoregions of the
United States (pp. 143-170). (NRS-80). Newtown Square, PA: U.S. Department of
Agriculture, Forest Service, Northern Research Station.
http://www.nrs.fs.fed.us/pubs/38109.
Fenn. ME: Allen. EB: Weiss. SB: Jovan. S: Geiser. LH: Tonnesen. GS: Johnson. RE: Rao. LE;
Gimeno. BS: Yuan. F; Meixner. T; Bytnerowicz. A. (2010). Nitrogen critical loads and
management alternatives for N-impacted ecosystems in California [Review], J Environ
Manage 91: 2404-2423. http://dx.doi.Org/10.1016/i.ienvman.2010.07.034.
Fenn. ME: Baron. JS: Allen. EB: Rueth. HM; Nvdick. KR: Geiser. L: Bowman. WD: Sickman.
JO: Meixner. T; Johnson. DW: Neitlich. P. (2003a). Ecological effects of nitrogen
deposition in the western United States. Bioscience 53: 404-420.
Fenn. ME: Driscoll. CT: Zhou. Q: Rao. LE: Meixner. T; Allen. EB: Yuan. F; Sullivan. TJ.
(2015). Use of Combined Biogeochemical Model Approaches and Empirical Data to
Assess Critical Loads of Nitrogen. In Environmental Pollution Series.
http://dx.doi.org/10.1007/978-94-017-95Q8-l 10.
Fenn. ME: Geiser. L: Bachman. R: Blubaugh. TJ: Bytnerowicz. A. (2007). Atmospheric
deposition inputs and effects on lichen chemistry and indicator species in the Columbia
River Gorge, USA. Environ Pollut 146: 77-91.
http://dx.doi.Org/10.1016/i.envpol.2006.06.024.
Fenn. ME: Geiser. LH. (2011). Temperate sierras. In Assessment of Nitrogen Deposition Effects
and Empirical Critical Loads of Nitrogen for Ecoregions of the United States (pp. 175-
180). (NRS-80). Newtown Square, PA: U.S. Department of Agriculture, Forest Service,
Northern Research Station, https://www.nrs.fs.fed.us/pubs/gtr/gtr nrs80.pdf.
Fenn. ME: Haeuber. R: Tonnesen. GS: Baron. JS: Grossman-Clarke. S: Hope. D: Jaffe. DA:
Copeland. S: Geiser. L; Rueth. HM: Sickman. JO. (2003b). Nitrogen emissions,
deposition, and monitoring in the western United States. Bioscience 53: 391-403.
Fenn. ME: Huntington. TG: McLaughlin. SB: Eagar. C: Gomez. A: Cook. RB. (2006). Status of
soil acidification in North America. J Forest Sci 52: 3-13.
Fenn, ME: Jovan, S; Yuan, F; Geiser, L; Meixner, T; Gimeno, BS. (2008). Empirical and
simulated critical loads for nitrogen deposition in California mixed conifer forests.
Environ Pollut 155: 492-511. http://dx.doi.Org/10.1016/i.envpol.2008.03.019.
Fenn, ME: Lambert, KF: Blett, TF: Burns, DA: Pardo, LH: Lovett, GM: Haeuber, RA: Evers,
DC: Driscoll, CT: Jefferies, DS. (201 lb). Setting limits: Using air pollution thresholds to
protect and restore U.S. ecosystems. (Issues in Ecology No. 14). Washington, DC:
Ecological Society of America, https://www.esa.org/esa/wp-
content/uploads/2013/03/issuesinecologyl4.pdf.
June 2018
R-69
DRAFT: Do Not Cite or Quote

-------
Fenn. ME; Poth. MA. (1999). Temporal and spatial trends in streamwater nitrate concentrations
in the San Bernardino Mountains, southern California. J Environ Qual 28: 822-836.
http://dx.doi.org/10.2134/ieql999.00472425002800030Q13x.
Fenn. ME: Poth. MA: Aber. JD; Baron. JS: Bonn arm. BT; Johnson. DW: Lemlv. AD: Mcnultv.
SG: Ryan. DF; Stottlemyer. R. (1998). Nitrogen excess in North American ecosystems:
Predisposing factors, ecosystem responses, and management strategies. Ecol Appl 8: 706-
733. http://dx.doi.org/10.1890/1051-0761(1998)008r0706:NEINAE12.0.CO;2.
Fenn. ME: Poth. MA: Bytnerowicz. A: Sickman. JO: Takemoto. BK. (2003c). Effects of ozone,
nitrogen deposition, and other stressors on montane ecosystems in the Sierra Nevada. In
A Bytnerowicz; MJ Arbaugh; R Alonso (Eds.), Developments in Environmental Science
(pp. 111-155). New York, NY: Elsevier. http://dx.doi.org/10.lQ16/S1474-
8177(02)02005-3.
Fenn. ME; Poth. MA; Johnson. DW. (1996). Evidence for nitrogen saturation in the San
Bernardino Mountains in southern California. For Ecol Manage 82: 211-230.
http://dx.doi.org/10.1016/0378-1127(95)03668-7.
Fenn. ME; Poth. MA; Terry. JD; Blubaugh. TJ. (2005). Nitrogen mineralization and nitrification
in a mixed-conifer forest in southern California: controlling factors, fluxes, and nitrogen
fertilization response at a high and low nitrogen deposition site. Can J For Res 35: 1464-
1486. http://dx.doi.org/10.1139/X05-068.
Fennel. K; Hetland. R; Feng. Y; Dimarco. S. (2011). A coupled physical-biological model of the
Northern Gulf of Mexico shelf: model description, validation and analysis of
phytoplankton variability. Biogeosciences 8: 1881-1899. http://dx.doi.org/10.5194/bg-8-
1881-2011.
Fernandes. I; Seena. S; Pascoal. C; Cassio. F. (2014). Elevated temperature may intensify the
positive effects of nutrients on microbial decomposition in streams. Freshw Biol 59:
2390-2399. http://dx.doi.org/10.1111/fwb. 12445.
Fernandes. MB; Benger. S; Stuart-Williams. H; Gavlard. S. am; Bryars. S. (2015). Coastal
nitrogen plumes and their relationship with seagrass distribution. Estuar Coast Shelf Sci
167: 390-403. http://dx.doi.Org/10.1016/i.ecss.2015.10.016.
Fernandez. IJ; Adams. MB; SanClements. MP; Norton. SA. (2010). Comparing decadal
responses of whole-water shed manipulations at the Bear Brook and Fernow experiments.
Environ Monit Assess 171: 149-161. http://dx.doi.org/10.1007/slQ661-010-1524-2.
Fernandez, IJ; Norton, SA. (2010). The Bear Brook Watershed in Maine—the second decade.
Preface. Environ Monit Assess 171: 1-2. http://dx.doi.org/10.1007/slQ661-010-1530-4.
Fernandez, IJ; Rustad, LE; Norton, SA; Kahl, JS; Cosby, BJ. (2003). Experimental acidification
causes soil base-cation depletion at the Bear Brook Watershed in Maine. Soil Sci Soc Am
J 67: 1909-1919. http://dx.doi.org/10.2136/sssai2003.1909.
Ferreira, JG; Bricker, SB. (2016). Goods and services of extensive aquaculture: shellfish culture
and nutrient trading. Aquaculture International 24: 803-825.
http://dx.doi.org/10.1007/slQ499-015-9949-9.
June 2018
R-70
DRAFT: Do Not Cite or Quote

-------
Ferreira. V; Castagnevrol, B; Koricheva. J; Gulis. V; Chauvet E; Graca. MAS. (2015). A meta-
analysis of the effects of nutrient enrichment on litter decomposition in streams [Review],
Biol Rev Camb Philos Soc 90: 669-688. http://dx.doi.ore/10.Ill 1/brv. 12125.
Ferretti. M; Marchetto. A: Arisci. S: Bussotti. F; Calderisi. M: Carnicelli. S: Cecchini. G: Fabbio.
G: Bertini. G: Matteucci. G: de Cinti. B: Salvati. L: Pompei. E. (2014). On the tracks of
Nitrogen deposition effects on temperate forests at their southern European range - an
observational study from Italy. Global Change Biol 20: 3423-3438.
http://dx.doi.org/10. Ill 1/gcb. 12552.
Fertig. B; Kennish. MJ; Sakowicz. GP: Reynolds. LK. (2014). Mind the data gap: Identifying
and assessing drivers of changing eutrophication condition. Estuaries Coasts 37: SI 98-
S221. http://dx.doi.org/10.1007/s 12237-013-9746-5.
Feuchtmavr. H; Moran. R; Hatton. K: Connor. L; Heves. T; Moss. B: Harvey. I: Atkinson. D.
(2009). Global warming and eutrophi cation: Effects on water chemistry and autotrophic
communities in experimental hypertrophic shallow lake mesocosms. J Appl Ecol 46: 713-
723. http://dx.doi.org/10.111 l/i l365-2664.2009.01644.x.
Field. CD: Pise. NB; Payne. RJ: Britton. AJ; Emmett. BA; Helliwell. RC: Hughes. S: Jones. L;
Lees. S: Leake. JR; Leith. I; Phoenix. GK; Power. SA; Sheppard. LJ; Southon. GE;
Stevens. CJ: Caporn. SJM. (2014). The role of nitrogen deposition in widespread plant
community change across semi-natural habitats. Ecosystems 17: 864-877.
http://dx.doi.org/10.1007/slQ021-014-9765-5.
Field. CD: Evans. CD: Pise. NB: Hall. JR: Caporn. SJ. (2017). Long-term nitrogen deposition
increases heathland carbon sequestration. Sci Total Environ 592: 426-435.
http://dx.doi.Org/10.1016/i.scitotenv.2017.03.059.
Fierer. N: Lauber. CL; Ramirez. KS: Zaneveld. J: Bradford. MA: Knight. R. (2012).
Comparative metagenomic, phylogenetic and physiological analyses of soil microbial
communities across nitrogen gradients. ISME J 6: 1007-1017.
http://dx.doi.org/10.1038/ismei.2011.159.
Findlav. PL. (2003). Response of phytoplankton communities to acidification and recovery in
Killarney Park and the Experimental Lakes Area, Ontario. Ambio 32: 190-195.
http://dx.doi.Org/10.1579/0044-7447-32.3.190.
Findlav. PL: Hecky. RE: Kasian. SEM; Stainton. MP: Hendzel. LL; Schindler. EU. (1999).
Effects on phytoplankton of nutrients added in conjunction with acidification. Freshw
Biol 41: 131-145. http://dx.doi.Org/10.1046/i.1365-2427.1999.00380.x.
Findlav. PL: Kasian. SEM. (1996). The effect of incremental pH recovery on the Lake 223
phytoplankton community. Can J Fish Aquat Sci 53: 856-864.
http://dx.doi.Org/10.l 139/f95-251.
Finlay, JC; Small, GE; Sterner, RW. (2013). Human influences on nitrogen removal in lakes.
Science 342: 247-250. http://dx.doi.Org/10.l 126/science. 1242575.
Finlev, MA: Courtenav, SC: Teather, KL: Hewitt, LM: Holdwav, PA: Hogan, NS: van Pen
Heuvel, MR. (2013). Evaluating cumulative effects of anthropogenic inputs in Prince
Edward Island estuaries using the mummichog (Fundulus heteroclitus). Integr Environ
Assess Manag 9: 496-507. http://dx.doi.org/10.1002/ieam.1396.
June 2018
R-71
DRAFT: Do Not Cite or Quote

-------
Finzi. AC. (2009). Decades of atmospheric deposition have not resulted in widespread
phosphorus limitation or saturation of tree demand for nitrogen in southern New England.
Biogeochemistry 92: 217-229. http://dx.doi.ore/10.1007/sl0533-009-9286-z.
Fischer. S: Frommen, JG. (2013). Eutrophication alters social preferences in three-spined
sticklebacks (Gasterosteus aculeatus). BehavEcol Sociobiol 67: 293-299.
http://dx.doi.org/10.1007/sQ0265-012-1449-6.
Fisher. JA; Jacob. DJ: Travis. KR; Kim. PS: Marais, EA; Chan miller. C: Yu, K; Zhu, L;
Yantosca. RM; Sulprizio. MP: Mao. J: Wennberg, PO; Crounse. JD; Teng, AP; Nguyen.
TB; St. Clair. JM; Cohen. RC: Romer. P; Nault. BA; Wooldridge. PJ: Jimenez. JL;
Campuzano-Jost. P; Day. DA: Shepson. PB; Xiong. F; Blake. PR: Goldstein. AH:
Misztal, PK: Hanisco, TF: Wolfe, GM: Ryerson, TB: Wisthaler, A: Mikoviny, T. (2016).
Organic nitrate chemistry and its implications for nitrogen budgets in an isoprene- and
monoterpene-rich atmosphere: Constraints from aircraft (SEAC4RS) and ground-based
(SOAS) observations in the Southeast US. Atmos Chem Phys Discuss 16: 5969-5991.
http://dx.doi.org/10.5194/acp-16-5969-2016.
Fisher. TR: Hagy III. JD: Boynton. WR: Williams. MR. (2006). Cultural eutrophication in the
Choptank and Patuxent estuaries of Chesapeake Bay. Limnology 51: 435-447.
http://dx.doi.Org/10.4319/lo.2006.51.l part 2.0435.
Fisher, TR; Lee, K, -Y; Berndt, H; Benitez, JA; Norton, MM. (1998). Hydrology and chemistry
of the Choptank River basin. Water Air Soil Pollut 105: 387-397.
http://dx.doi.Org/10.1023/A:1005001525063.
Fisk, MC: Schmidt, SK: Seastedt, TR. (1998). Topographic patterns of above and below ground
production and nitrogen cycling in Alpine tundra. Ecology 79: 2253-2266.
http://dx.doi.org/10.1890/0012-9658(1998N)079r2253:TPOAAB12.0.CQ:2.
Fismes, J: Perrin-Ganier, C: Empereur-Bissonnet, P; Morel, JL. (2002). Soil-to-root transfer and
translocation of polycyclic aromatic hydrocarbons by vegetables grown on industrial
contaminated soils. J Environ Qual 31: 1649-1656.
Fiellheim, A: Raddum, GG: Sagen, T. (1985). Effect of aluminum at low pH on the mortality of
elvers (Anguilla anguilla L.), a laboratory experiment. J Exp Zool Part A Ecol Genet
Physiol 22: 2544-2547.
Flechard, CR: Fowler, D: Sutton, MA: Cape, JN. (1999). A dynamic chemical model of bi-
directional ammonia exchange between semi-natural vegetation and the atmosphere. Q J
Roy Meteorol Soc 125: 2611-2641.
Flechard, CR: Massad, RS: Loubet, B: Personne, E: Simpson, D: Bash, JO: Cooter, EJ: Nemitz,
E: Sutton, MA. (2013). Advances in understanding, models and parameterizations of
biosphere-atmosphere ammonia exchange. Biogeosciences 10: 5183-5225.
http://dx.doi.org/10.5194/bg-10-5183-2013.
Fleischer, K: Rebel, KT: van der Molen, MK: Erisman, JW: Wassen, MJ: van Loon, EE:
Montagnani, L: Gough, CM: Herbst, M: Janssens, IA: Gianelle, D: Dolman, AJ. (2013).
The contribution of nitrogen deposition to the photosynthetic capacity of forests. Global
Biogeochem Cycles 27: 187-199. http://dx.doi.org/10.1002/gbc.20Q26.
June 2018
R-72
DRAFT: Do Not Cite or Quote

-------
Florida DEP. Estuary-specific numeric interpretations of the narrative nutrient criterion. (Florida
Aministrative Code2016). https://www.flrules.org/gatewav/ruleno.asp?id=62-302.532
Flum, T; Nodvin, SC. (1995). Factors affecting streamwater chemistry in the Great Smoky
Mountains, USA. Water Air Soil Pollut 85: 1707-1712.
http://dx.doi.org/10.1007/BF0Q477226.
Ford. J: Stoddard. JL; Powers. CF. (1993). Perspectives on environmental monitoring: An
introduction to the US EPA Long-Term Monitonring (LTM) project. Water Air Soil
Pollut 67: 247-255. http://dx.doi.org/10.1007/BF0Q478148.
Fork, ML; Heffernan, JB. (2014). Direct and indirect effects of dissolved organic matter source
and concentration on denitrification in Northern Florida rivers. Ecosystems 17: 14-28.
http://dx.doi.org/10.1007/slQ021-013-9705-9.
Forsius. M; Posch. M; Aherne. J: Reinds. GJ: Christensen. J: Hole. L. (2010). Assessing the
impacts of long-range sulfur and nitrogen deposition on arctic and sub-arctic ecosystems.
Ambio 39: 136-147. http://dx.doi.org/10.1007/sl3280-010-0Q22-7.
Fowler. D; Covle. M; Skiba. U: Sutton. MA: Cape. JN; Reis. S: Sheppard. LJ; Jenkins. A:
Grizzetti. B; Galloway. JN; Vitousek. P; Leach. A: Bouwman. AF; Butterbach-Bahl. K;
Dentener, F; Stevenson, D; Amann, M; Yoss, M. (2013). The global nitrogen cycle in the
twenty-first century [Review], Philos Trans R Soc Lond B Biol Sci 368: 20130164.
http://dx.doi.org/10.1098/rstb.2013.0164.
Fowler, D; Pilegaard, K; Sutton, MA: Ambus, P; Raivonen, M; Duvzer, J: Simpson, D; Fagerli,
Ft; Fuzzi, S; Schjoerring, JK; Granier, C; Neftel, A; Isaksen, ISA; Lai, P; Mai one, M;
Monks, PS; Burkhardt, J; Daemmgen, U; Neirynck, J; Personne, E; Wichink-Kruit, R;
Butterbach-Bahl, K; Flechard, C; Tuovinen, JP; Covle, M; Gerosa, G; Loubet, B; Altimir,
N; Gruenhage, L; Ammann, C; Cieslik, S; Paoletti, E; Mikkelsen, TN; Ro-Poulsen, H;
Cellier, P; Cape, JN; Horvath, L; Loreto, F; Niinemets, U; Palmer, PI; Rinne, J; Misztal,
P; Nemitz, E; Nilsson, D; Pryor, S; Gallagher, MW; Vesala, T; Skiba, U; Briiggemann,
N; Zechmeister-Boltenstern, S; Williams, J; O'Dowd, C; Facchini, MC; de Leeuw, G;
Flossman, A; Chaumerliac, N; Erisman, JW. (2009). Atmospheric composition change:
Ecosystems-atmosphere interactions. Atmos Environ 43: 5193-5267.
http://dx.doi.Org/10.1016/i.atmosenv.2009.07.068.
Fowler, ZK; Adams, MB; Peterjohn, WT. (2015). Will more nitrogen enhance carbon storage in
young forest stands in central Appalachia? For Ecol Manage 337: 144-152.
http://dx.doi.Org/10.1016/i.foreco.2014.10.023.
Fox, L; Valiela, I; Kinney, EL. (2012). Vegetation Cover and Elevation in Long-Term
Experimental Nutrient-Enrichment Plots in Great Sippewissett Salt Marsh, Cape Cod,
Massachusetts: Implications for Eutrophication and Sea Level rise. Estuaries Coasts 35:
445-458. http://dx.doi.org/10.1007/sl2237-012-9479-x.
Frame, JL; Jones, JI; Ormerod, SJ; Sadler, JP; Ledger, ME. (2016). Biological barriers to
restoration: testing the biotic resistance hypothesis in an upland stream recovering from
acidification. Hydrobiologia 777: 161-170. http://dx.doi.org/10.1007/slQ750-016-2772-0.
June 2018
R-73
DRAFT: Do Not Cite or Quote

-------
Fraser. C; Aim. EJ; Polz. MF; Spratt BG; Hanage, WP. (2009). The bacterial species challenge:
Making sense of genetic and ecological diversity [Review], Science 323: 741-746.
http://dx.doi.ore/10.1126/science.l 159388.
Freda. J: Dunson. WA. (1986). Effects of low pH and other chemical variables on the local
distribution of amphibians. Copeia 1986: 454-466. http://dx.doi.org/10.2307/1445003.
Freedman. Z; Eisenlord. SD; Zak. PR: Xue. K; He. Z; Zhou. J. (2013). Towards a molecular
understanding of N cycling in northern hardwood forests under future rates of N
deposition. Soil Biol Biochem 66: 130-138.
http://dx.doi.Org/10.1016/i.soilbio.2013.07.010.
Freedman. Z; Zak. DR. (2014). Atmospheric N deposition increases bacterial laccase-like
multicopper oxidases: Implications for organic matter decay. Appl Environ Microbiol 80:
4460-4468. http://dx.doi.org/10.1128/AEM.01224-14.
Freedman, ZB; Upchurch, RA; Zak, PR; Cline, LC. (2016). Anthropogenic N deposition slows
decay by favoring bacterial metabolism: Insights from metagenomic analyses. FMICB 7:
1-11. http://dx.doi.org/10.3389/fmicb.2016.00259.
Frenette, J, -J: Richard, Y; Moreau, G. (1986). Fish responses to acidity in Quebec lakes: A
review [Review], Water Air Soil Pollut 30: 461-475.
http://dx.doi.org/10.1007/BF003Q5215.
Frev, SD: Knorr, M; Parrent, JL; Simpson, RT. (2004). Chronic nitrogen enrichment affects the
structure and function of the soil microbial community in temperate hardwood and pine
forests. For Ecol Manage 196: 159-171. http://dx.doi.Org/10.1016/i.foreco.2004.03.018.
Frev, SD: Ollinger, S: Nadelhoffer, K; Bowden, R; Brzostek, E; Burton, A: Caldwell, BA; Crow,
S: Goodale, CL; Grandv, AS: Finzi, AC: Kramer, MG: Laitha, K; Lemoine, J: Martin, M;
Mcdowell, WH: Minocha, R; Sadowsky, JJ; Templer, PH; Wickings, K. (2014). Chronic
nitrogen additions suppress decomposition and sequester soil carbon in temperate forests.
Biogeochemistry 121: 305-316. http://dx.doi.org/10.1007/sl0533-014-0004-Q.
Friberg, N: Skriver, J: Larsen, SE; Pedersen, ML: Buffagni, A. (2010). Stream macroinvertebrate
occurrence along gradients in organic pollution and eutrophication. Freshw Biol 55:
1405-1419. http://dx.doi.org/10.111 l/j.l365-2427.2008.02164.x.
Friedrich, U; von Oheimb, G: Kriebitzsch, W, -U: Schlefielmann, K; Weber, MS: Hardtle, W.
(2012). Nitrogen deposition increases susceptibility to drought - experimental evidence
with the perennial grass Molinia caerulea (L.) Moench. Plant Soil 353: 59-71.
http://dx.doi.org/10.1007/sl 1104-011-1008-3.
Frindte, K; Allgaier, M; Grossart, H, -P; Eckert, W. (2015). Microbial response to
experimentally controlled redox transitions at the sediment water interface. PLoS ONE
10: e0143428. http://dx.doi.org/10.1371/iournal.pone.0143428.
Fritz, C: Lamers, LPM; Riaz, M; van den Berg, LJL; Elzenga, TJT, M. (2014). Sphagnum
mosses - masters of efficient N-uptake while avoiding Intoxication. PLoS ONE 9:
e79991. http://dx.doi.org/10.1371/iournal.pone.0079991.
June 2018
R-74
DRAFT: Do Not Cite or Quote

-------
Frohne. T; Rinklebe. J; Langer, U: Du Laing, G; Mothes. S; Wennrich. R. (2012).
Biogeochemical factors affecting mercury methylation rate in two contaminated
floodplain soils. Biogeosciences 9: 493-507. http ://dx.doi.org/10.5194/bg-9-493-2012.
From. F; Strengbom. J; Nordin. A. (2015). Residual long-term effects of forest fertilization on
tree growth and nitrogen turnover in boreal forest. 6: 1145-1156.
http://dx.doi.org/10.3390/f6Q41145.
Frost JW; Schleicher. T; Craft C. (2009). Effects of nitrogen and phosphorus additions on
primary production and invertebrate densities in a Georgia (USA) tidal freshwater marsh.
Wetlands 29: 196-203. http://dx.doi.Org/10.1672/07-79.l.
Frost. TM; Fischer. JM; Klug. JL; Arnott. SE; Montz. PK. (2006). Trajectories of zooplankton
recovery in the Little Rock Lake whole-lake acidification experiment. Ecol Appl 16: 353-
367. http://dx.doi.org/10.1890/04-180Q.
Frost, TM; Montz, PK; Kratz, TK; Badillo, T; Brezonik, P; Gonzalez, MJ; Rada, RG; Watras,
CJ; Webster, KE; Wiener, JG; Williamson, CE; Morris, DP. (1999). Multiple stresses
from a single agent: Diverse responses to the experimental acidification of Little Rock
Lake, Wisconsin. Limnol Oceanogr 44: 784-794.
http://dx.doi.Org/10.4319/lo.1999.44.3 part 2.0784.
Fu, G; Shen, ZX, i. (2016). Response of Alpine Plants to Nitrogen Addition on the Tibetan
Plateau: A Meta-analysis. Journal of Plant Growth Regulation 35: 974-979.
http://dx.doi.org/10.1007/s00344-016-9595-Q.
Fu, G; Shen, ZX, i. (2017). Response of alpine soils to nitrogen addition on the Tibetan Plateau:
A meta-analysis. Appl Soil Ecol 114: 99-104.
http://dx.doi.Org/10.1016/i.apsoil.2017.03.008.
Fulweiler, RW; Heiss, EM. (2014). (Nearly) A Decade of Directly Measured Sediment N-2
Fluxes What Can Narragansett Bay Tell Us About the Global Ocean Nitrogen Budget?
Oceanography 27: 184-195.
Fulweiler, RW; Nixon, SW; Buckley, BA; Granger, SL. (2008). Net Sediment N-2 Fluxes in a
Coastal Marine System-Experimental Manipulations and a Conceptual Model.
Ecosystems 11: 1168-1180. http://dx.doi.org/10.1007/slQ021-008-9187-3.
Fusaro, L; Salvatori, E; Manes, F. (2017). Effects of nitrogen deposition, drought and their
interaction, on functional and structural traits of Fraxinus ornus L. and Quercus ilex L.
Plant Biosyst 151: 174-189. http://dx.doi.org/10.1080/11263504.2016.119307Q.
Fuss, CB; Driscoll, CT; Campbell, JL. (2015). Recovery from chronic and snowmelt
acidification: Long-term trends in stream and soil water chemistry at the Hubbard Brook
Experimental Forest, New Hampshire, USA. Jour Geo Res: Biog 120: 2360-2374.
http://dx.doi. org/10.1002/2015JG003 063.
Futter, MN; Klaminder, J; Lucas, RW; Laudon, H; Kohler, SJ. (2012). Uncertainty in silicate
mineral weathering rate estimates: Source partitioning and policy implications. Environ
Res Lett 7: 1-8. http://dx.doi.Org/10.1088/1748-9326/7/2/024025.
G.Nelson, W. (2017). Development of an epiphyte indicator of nutrient enrichment: Threshold
values for seagrass epiphyte load. Ecol Indicat 74: 343-356.
June 2018
R-75
DRAFT: Do Not Cite or Quote

-------
Gabriel MC; Axelrad. D; Orem. W; Osborne. TZ. (2015). Response to Julian et al. (2015)
"Comment on and Reinterpretation of Gabriel et al. (2014) 'Fish Mercury and Surface
Water Sulfate Relationships in the Everglades Protection Area'" [Comment], Environ
Manage 55: 1227-1231. http://dx.doi.org/10.1007/s00267-015-0486-Q.
Gabriel MC: Howard. N: Osborne. TZ. (2014). Fish mercury and surface water sulfate
relationships in the Everglades protection area. Environ Manage 53: 583-593.
http://dx.doi. org/10.1007/s00267-013 -0224-4.
Gaio-Oliveira. G: Dahlman. L; Palmqvist. K; Martins-Loucao. MA: Maguas. C. (2005). Nitrogen
uptake in relation to excess supply and its effects on the lichens Evernia prunastri (L.)
Ach and Xanthoria parietina (L.) Th. Fr. Planta 220: 794-803.
http://dx.doi.org/10.1007/sQ0425-004-1396-l.
Gall. JE; Boyd. RS: Raiakaruna. N. (2015). Transfer of heavy metals through terrestrial food
webs: a review. Environ Monit Assess 187. http://dx.doi.org/10.1007/slQ661-015-4436-
3.
Gallagher. MW; Beswick. KM: Duvzer. J: Westrate. H: Choularton. TW: Hummelsh0i. P.
(1997). Measurements of aerosol fluxes to Speulder forest using a micrometeorological
technique. Atmos Environ 31: 359-373. http://dx.doi.Org/10.1016/S 1352-2310(96)00057-
X.
Gallagher. MW: Choularton. TW: Morse. AP; Fowler. D. (1988). Measurements of the size
dependence of cloud droplet deposition at a hill site. Q J Roy Meteorol Soc 114: 1291-
1303. http://dx.doi.org/10.1002/ai.49711448307.
Galloway. JN. (1998). The global nitrogen cycle: Changes and consequences. Environ Pollut 1:
15-24. http://dx.doi.org/10.1016/50269-7491(98)80010-9.
Galloway. JN: Aber. JD; Erisman. JW: Seitzinger. SP; Howarth. RW: Cowling. EB; Cosby. BJ.
(2003). The nitrogen cascade. Bioscience 53: 341-356.
Galloway. JN: Cowling. EB. (2002). Reactive nitrogen and the world: 200 years of change
[Review], Ambio 31: 64-71. http://dx.doi.org/10.1639/0Q44-
7447(2002)031 r0064:RNATWY12.0.CQ:2.
Galloway. JN: Dentener. FJ: Capone. DG: Bover. EW: Howarth. RW: Seotzomger Seitzinger.
SP: Asner. GP: Cleveland. CC: Green. PA: Holland. EA: Karl. DM: Michaels. AF;
Porter. JH: Townsend. AR: Vorosmartv. CJ. (2004). Nitrogen cycles: Past, present, and
future. Biogeochemistry 70: 153-226. http://dx.doi.org/10.1007/sl0533-004-0370-Q.
Galloway. JN: Likens. GE: Edgerton. ES. (1976). Acid precipitation in the Northeastern United
States: pH and acidity. Science 194: 722-724.
http://dx.doi.Org/10.l 126/science. 194.4266.722.
Galloway. JN: Likens. GE: Keene. WC: Miller. JM. (1982). The composition of precipitation in
remote areas of the world. J Geophys Res 87: 8771-8786.
http://dx.doi.org/10.1029/JC087iCllp08771.
June 2018
R-76
DRAFT: Do Not Cite or Quote

-------
Galloway. JN; Schlesinger. WH; Clark. CM; Grimm. NB; Jackson. RB; Law. BE; Thornton. PE;
Townsend. AR; Martin. R. (2014). Ch. 15: Biogeochemical cycles. In JM Melillo; TC
Richmond; GW Yohe (Eds.), Climate change impacts in the United States: The third
National Climate Assessment (pp. 350-368). Washington, DC: U.S. Global Change
Research Program. http://dx.doi.org/10.7930/J0X63JT0.
Galloway. JN; Schlesinger. WH; Levy II. H; Michaels. A; Schnoor. JL. (1995). Nitrogen-
fixation: Anthropogenic enhancement-environmental response. Global Biogeochem
Cycles 9: 235-252. http://dx.doi.org/10.1029/95GB00158.
Galloway. JN; Townsend. AR; Erisman. JW; Bekunda. M; Cai. ZC; Frenev. JR; Martinelli. LA;
Seitzinger. SP; Sutton. MA. (2008). Transformation of the nitrogen cycle: Recent trends,
questions, and potential solutions [Review], Science 320: 889-892.
http://dx.doi.Org/10.l 126/science.l 136674.
Gamalero. E; Lingua. G; Berta. G; Glick. BR. (2009). Beneficial role of plant growth promoting
bacteria and arbuscular mycorrhizal fungi on plant responses to heavy metal stress. Can J
Microbiol 55: 501-514. http://dx.doi.org/10.1139/W09-010.
Gaman. A; Rannik. U; Aalto. P; Pohia. T; Siivola. E; Kulmala. M; Vesala. T. (2004). Relaxed
eddy accumulation system for size-resolved aerosol particle flux measurements. J Atmos
Ocean Tech 21: 933-943. http://dx.doi.org/10.1175/152Q-
0426(2004)021<0933:REASFS>2.0.CQ;2.
Gamfeldt. L; Lefcheck. JS; Byrnes. JEK; Cardinale. BJ; Duffy. JE; Griffin. JN. (2015). Marine
biodiversity and ecosystem functioning: what's known and what's next? Oikos 124: 252-
265. http://dx.doi.org/10. Ill l/oik.01549.
Gamfeldt. L; Snail. T; Bagchi. R; Jonsson. M; Gustafsson. L; Kiellander. P; Ruiz-Jaen. MC;
Froberg. M; Stendahl. J; Philipson. CD; Mikusinski. G; Andersson. E; Westerlund. B;
Andren. H; Moberg. F; Moen. J; Bengtsson. J. (2013). Higher levels of multiple
ecosystem services are found in forests with more tree species. Nat Commun 4: 1340.
http://dx.doi.org/10.1038/ncomms2328.
Gan. H; Zak. PR; Hunter. MP. (2013). Chronic nitrogen deposition alters the structure and
function of detrital food webs in a northern hardwood ecosystem. Ecol Appl 23: 1311-
1321. http://dx.doi.Org/10.1890/12-1895.l.
Gan. H; Zak. PR; Hunter. MP. (2014). Trophic stability of soil oribatid mites in the face of
environmental change. Soil Biol Biochem 68: 71-77.
http://dx.doi.Org/10.1016/i.soilbio.2013.09.019.
Ganzeveld. L; Lelieveld. J; Pentener. FJ; Krol. MC; Bouwman. AJ; Roelofs. GJ. (2002). Global
soil-biogenic NOx emissions and the role of canopy processes. J Geophys Res 107.
Gao. G; Clare. AS; Rose. C; Caldwell. GS. (2017). Eutrophication and warming-driven green
tides (Ulva rigida) are predicted to increase under future climate change scenarios. Mar
PollutBull 114: 439-447. http://dx.doi.Org/10.1016/i.marpolbul.2016.10.003.
Garcia. MO; Ovasapyan. T; Greas. M; Treseder. KK. (2008). Mycorrhizal dynamics under
elevated C02 and nitrogen fertilization in a warm temperate forest. Plant Soil 303: 301-
310. http://dx.doi.org/10.1007/slll04-007-95Q9-9.
June 2018
R-77
DRAFT: Do Not Cite or Quote

-------
Gardner. EM; McKnight, DM; Lewis. WM, Jr; Miller. MP. (2008). Effects of nutrient
enrichment on phytoplankton in an Alpine Lake, Colorado, U.S.A. Arctic Antarct Alpine
Res 40: 55-64. http://dx.doi.org/10.1657/1523-0430(07-002)rGARPNER12.0.CQ;2.
Gardner. WS: Seitzinger. SP: Malczvk. JM. (1991). THE EFFECTS OF SEA SALTS ON THE
FORMS OF NITROGEN RELEASED FROM ESTUARINE AND FRESH-WATER
SEDIMENTS - DOES ION-PAIRING AFFECT AMMONIUM FLUX. Estuaries 14:
157-166.
Garmendia. M; Bricker. S: Revilla. M; Boria. A; Franco. J: Bald. J: Valencia. V. (2012).
Eutrophication assessment in Basque estuaries: Comparing a North American and a
European method. Estuaries Coasts 35: 991-1006. http://dx.doi.org/10.1007/sl2237-Q12-
9489-8.
Garten. CT. Jr. (2000). Nitrogen saturation and soil N availability in a high-elevation spruce and
fir forest. Water Air Soil Pollut 120: 295-313.
http://dx.doi.Org/10.1023/A:1005270012802.
Gascon Diez. E; Loizeau. JL; Cosio. C: Bouchet. S: Adatte. T; Amouroux. D; Bravo. AG.
(2016). Role of Settling Particles on Mercury Methylation in the Oxic Water Column of
Freshwater Systems. Environ Sci Technol 50: 11672-11679.
http://dx.doi.org/10.1021/acs.est.6b03260.
Gassman. PW: Reyes. MR: Green. CH; Arnold. JG. (2007). The soil and water assessment tool:
Historical development, applications, and future research directions. Trans ASAE 50:
1211-1250.
Gates. DM. (1968). Transpiration and leaf temperature. Annu Rev Plant Biol 19: 211-238.
http://dx.doi.Org/10.l 146/annurev.pp. 19.060168.001235.
Gauci. V: Fowler. D; Chapman. SJ: Pise. NB. (2004). Sulfate deposition and temperature
controls on methane emission and sulfur forms in peat. Biogeochemistry 71: 141-162.
http://dx.doi.org/10.1007/slQ533-004-9681-4.
Gaudnik. C: Corcket. E; Clement. B; Delmas. CEL; Gombert-Courvoisier. S: Muller. S: Stevens.
CJ: Alard. D. (2011). Detecting the footprint of changing atmospheric nitrogen deposition
loads on acid grasslands in the context of climate change. Global Change Biol 17: 3351-
3365. http://dx.doi.org/10.1111/i. 1365-2486.2011,02463.x.
Gavlord. B; Kroeker. KJ; Sunday. JM: Anderson. KM: Barry. JP; Brown. NE; Connell. SD;
Dupont. S. am: Fabricius. KE; Hall-Spencer. JM: Klinger. T; Milazzo. M; Mundav. PL:
Russell, BP; Sanford, E; Schreiber, SJ: Thiyagarai an, V: Vaughan, MLH; Widdicombe,
S: Harlev, CDG. (2015). Ocean acidification through the lens of ecological theory.
Ecology 96: 3-15. http://dx.doi.Org/10.1890/14-0802.l.
Gbondo-Tugbawa, SS: Priscoll, CT. (2002). Evaluation of the effects of future controls on sulfur
dioxide and nitrogen oxide emissions on the acid-base status of a northern forest
ecosystem. Atmos Environ 36: 1631-1643. http://dx.doi.org/10.lQ16/S1352-
2310(02^)00082-1.
June 2018
R-78
DRAFT: Do Not Cite or Quote

-------
Gbondo-Tugbawa. SS; Driscoll. CT. (2003). Factors controlling long-term changes in soil pools
of exchangeable basic cations and stream acid neutralizing capacity in a northern
hardwood forest ecosystem. Biogeochemistry 63: 161-185.
http://dx.doi.org/10.1023/a: 1023308525316.
Gbondo-Tugbawa. SS: Driscoll CT: Aber. JD; Likens. GE. (2001). Evaluation of an integrated
biogeochemical model (PnET-BGC) at a northern hardwood forest ecosystem. Water
Resour Res 37: 1057-1070. http://dx.doi.org/10.1029/2000WR90Q375.
Gebhart. KA; Schichtel. BA; Malm. WC: Barn a. MG: Rodriguez. MA: Collett JL. (2011). Back-
trajectory-based source apportionment of airborne sulfur and nitrogen concentrations at
Rocky Mountain National Park, Colorado, USA. Atmos Environ 45: 621-633.
http://dx.doi.Org/10.1016/i.atmosenv.2010.10.035.
Geiser. LH; Jovan. SE; Glavich. DA: Porter. MK. (2010). Lichen-based critical loads for
atmospheric nitrogen deposition in Western Oregon and Washington Forests, USA.
EnvironPollut 158: 2412-2421. http://dx.doi.Org/10.1016/i.envpol.2010.04.001.
Geiser. LH: Neitlich. PN. (2007). Air pollution and climate gradients in western Oregon and
Washington indicated by epiphytic macrolichens. Environ Pollut 145: 203-218.
http://dx.doi.Org/10.1016/i.envpol.2006.03.024.
Geisseler. D; Lazicki. PA: Scow. KM. (2016). Mineral nitrogen input decreases microbial
biomass in soils under grasslands but not annual crops. Appl Soil Ecol 106: 1-10.
http://dx.doi.Org/10.1016/i.apsoil.2016.04.015.
Gennaro, P; Piazzi, L; Persia, E; Porrello, S. (2015). Nutrient exploitation and competition
strategies of the invasive seaweed Caulerpa cylindracea. European Journal of Phycology
50: 384-394. http://dx.doi.org/10.1080/09670262.2Q15.1055591.
Gerber. AM: Ginn. BK: Whitfield. CJ: Dillon. PJ: Cumming. BF: Smol. JP. (2008). Glasgow
lake: An early-warning sentinel of lake acidification in cape breton highlands national
park (Nova scotia, Canada). Hydrobiologia 614: 299-307.
http://dx.doi.org/10.1007/slQ750-008-9514-x.
Gerber. S: Hedin. LO: Keel. SG: Pacala. SW: Shevliakova. E. (2013). Land use change and
nitrogen feedbacks constrain the trajectory of the land carbon sink. Geophys Res Lett 40:
5218-5222. http://dx.doi.org/10.1002/grl.50957.
Gerber. S: Hedin. LO: Oppenheimer. M: Pacala. SW: Shevliakova. E. (2010). Nitrogen cycling
and feedbacks in a global dynamic land model. Global Biogeochem Cycles 24.
http://dx.doi.org/10.1029/2008GB0Q3336.
Gerdol. R: Brancaleoni. L. (2015). Slow recovery of mire vegetation from environmental
perturbations caused by a heat wave and experimental fertilization. Wetlands 35: 769-
782. http://dx.doi.org/10.1007/sl3157-015-Q668-9.
Geron. C: Rasmussen. R: Arnts. RR: Guenther. A. (2000). A review and synthesis of
monoterpene speciation from forests in the United States [Review], Atmos Environ 34:
1761-1781. http://dx.doi.org/10.1016/S1352-2310(99^00364-7.
June 2018
R-79
DRAFT: Do Not Cite or Quote

-------
Gerson, JR.: Driscoll, CT. (2016). Is Mercury in a Remote Forested Watershed of the Adirondack
Mountains Responding to Recent Decreases in Emissions? Environ Sci Technol 50:
10943-10950. http://dx.doi.org/10.1021/acs.est.6b02127.
Gerson, JR: Driscoll. CT: Roy. KM. (2016). Patterns of nutrient dynamics in Adirondack lakes
recovering from acid deposition. Ecol Appl 26: 1758-1770. http://dx.doi.org/10.189Q/15-
1361.1.
Gibble, CM: Kudela, RM. (2014). Detection of persistent microcystin toxins at the land-sea
interface in Monterey Bay, California. Harmful Algae 39: 146-153.
http://dx.doi.Org/10.1016/i.hal.2014.07.004.
Giblin. AE: Tobias. CR: Song. B: Weston. N: Bant a. GT: Rivera-Monrov. VH. (2013). The
Importance of Dissimilatory Nitrate Reduction to Ammonium (DNRA) in the Nitrogen
Cycle of Coastal Ecosystems. Oceanography 26: 124-131.
Gibson, MP: Heal, MR: Li, Z: Kuchta, J: King, GH: Haves, A: Lambert, S. (2013). The spatial
and seasonal variation of nitrogen dioxide and sulfur dioxide in Cape Breton Highlands
National Park, Canada, and the association with lichen abundance. Atmos Environ 64:
303-311. http://dx.doi.Org/10.1016/i.atmosenv.2012.09.068.
Gielen, B: Calfapietra, C: Lukac, M: Wittig, YE: De Angelis, P: Janssens, IA: Moscatelli, MC:
Grego, S: Cotrufo, MF: Godbold, PL: Hoosbeek, MR: Long, SP: Miglietta, F: Polle, A:
Bernacchi, CJ: Davev, PA: Ceulemans, R: Scarascia-Mugnozza, GE. (2005). Net carbon
storage in a poplar plantation (POPFACE) after three years of free-air C02 enrichment.
Tree Physiol 25: 1399-1408. http://dx.doi.Org/10.1093/treephys/25.l 1.1399.
Gill, RA. (2014). The influence of 3-years of warming and N-deposition on ecosystem dynamics
is small compared to past land use in subalpine meadows. Plant Soil 374: 197-210.
http://dx.doi.org/10.1007/sl 1104-013-1868-9.
Gillet, F: Peter, M: Aver, F: Butler, R: Egli, S. (2010). Long-term dynamics of aboveground
fungal communities in a subalpine Norway spruce forest under elevated nitrogen input.
Oecologia 164: 499-510. http://dx.doi.org/10.1007/sQ0442-010-1668-3.
Gilliam, FS. (2006). Response of the herbaceous layer of forest ecosystems to excess nitrogen
deposition. J Ecol 94: 1176-1191. http://dx.doi.org/10.1111/j. 1365-2745.2006.01155.x.
Gilliam, FS. (2007). The ecological significance of the herbaceous layer in temperate forest
ecosystems. Bioscience 57: 845-858. http://dx.doi.org/10.1641/b571007.
Gilliam, FS. (2016). Forest ecosystems of temperate climatic regions: from ancient use to climate
change [Review], New Phytol 212: 871-887. http://dx.doi.org/10.Ill 1/nph. 14255.
Gilliam, FS: Adams, MB: Yurish, BM. (1996). Ecosystem nutrient responses to chronic nitrogen
inputs at Fernow Experimental Forest, West Virginia. Can J For Res 26: 196-205.
http://dx.doi.Org/10.l 139/x26-023.
Gilliam, FS: Billmyer, JH: Walter, CA: Peteriohn, WT. (2016a). Effects of excess nitrogen on
biogeochemistry of a temperate hardwood forest: Evidence of nutrient redistribution by a
forest understory species. Atmos Environ 146: 261-270.
http://dx.doi.Org/10.1016/i.atmosenv.2016.04.007.
June 2018
R-80
DRAFT: Do Not Cite or Quote

-------
Gilliam. FS; Goodale. CL; Pardo. LH; Geiser. LH; Lilleskov. EA. (201 la). Eastern temperate
forests. In Assessment of nitrogen deposition effects and empirical critical loads of
nitrogen for ecoregions of the United States (pp. 99-116). (NRS-80). Newtown Square,
PA: U.S. Department of Agriculture, Forest Service, Northern Research Station.
Gilliam. FS; Hockenberry. AW; Adams. MB. (2006). Effects of atmospheric nitrogen deposition
on the herbaceous layer of a central Appalachian hardwood forest. Torrey Botanical
Society Journal 133: 240-254. http://dx.doi.ore/10.3159/1095-
5674(2006N)133r240:EOAND012.0.CQ:2.
Gilliam. FS; McCullev. RL; Nelson. JA. (201 lb). Spatial variability in soil microbial
communities in a nitrogen-saturated hardwood forest watershed. Soil Sci Soc Am J 75:
280. http://dx.doi.ore/10.2136/sssai2010.0172.
Gilliam. FS; Turrill. NL; Aulick. SD; Evans. DK; Adams. MB. (1994). Herbaceous layer and
soil response to experimental acidification in a central appalachian hardwood forest. J
Environ Qual 23: 835-844. http://dx.doi.ore/10.2134/ieal994.00472425002300040032x.
Gilliam. FS; Welch. NT; Phillips. AH; Billmyer. JH; Peteriohn. WT; Fowler. ZK; Walter. CA;
Burnham. MB; May. JD; Adams. MB. (2016b). Twenty-five-year response of the
herbaceous layer of a temperate hardwood forest to elevated nitrogen deposition.
Ecosphere 7. http://dx.doi.ore/10.1002/ecs2.1250.
Gilliland. A; Dennis. R; Roselle. S; Pierce. T. (2003). Seasonal NH3 emission estimates for the
eastern United States based on ammonium wet concentrations and an inverse modeling
method. J Geophys Res 108: Article #4477. http://dx.doi.ore/10.1029/2002JD003063.
Gilmour. CC; Henry. EA; Mitchell. R. (1992). Sulfate stimulation of mercury methylation in
freshwater sediments. Environ Sci Technol 26: 2281-2287.
http://dx.doi.ore/10.1021/es00035a029.
Gilmour, CC; Podar, M; Bullock, AL; Graham, AM; Brown, SD; Somenahally, AC; Johs, A;
Hurt, RA, Jr; Bailey, KL; Elias, DA. (2013). Mercury methylation by novel
microorganisms from new environments. Environ Sci Technol 47: 11810-11820.
http://dx.doi.ore/10.1021/es403075t.
Gilmour, CC; Riedel, GS. (1995). Measurement of Hg methylation in sediments using high
specific-activity 203Hg and ambient incubation. Water Air Soil Pollut 80: 747-756.
http://dx.doi.ore/10.1007/BF01189726.
Gilmour, CC; Riedel, GS; Ederineton, MC; Bell, JT; Benoit, JM; Gill. GA; Stordal, MC. (1998).
Methylmercury concentrations and production rates across a trophic gradient in the
northern everglades. Biogeochemistry 40: 327-345.
http://dx.doi.ore/10.1023/A: 1005972708616.
Giedrem. T; Rosseland. BO. (2012). Genetic variation for tolerance to acidic water in salmonids
[Review], J Fish Biol 80: 1-14. http://dx.doi.ore/10.1111/i. 1095-8649.2011.03159.x.
Glassman. SI; Casper. BB. (2012). Biotic contexts alter metal sequestration and AMF effects on
plant growth in soils polluted with heavy metals. Ecology 93: 1550-1559.
http://dx.doi.ore/10.1890/10-213 5.1.
June 2018
R-81
DRAFT: Do Not Cite or Quote

-------
Glavich. DA; Geiser. LH. (2008). Potential approaches to developing lichen-based critical loads
and levels for nitrogen, sulfur and metal-containing atmospheric pollutants in North
America. Bryologist 111: 638-649. http://dx.doi.ore/10.1639/0007-2745-l 11.4.638.
Gledhill. DK; White. MM: Salisbury. J: Thomas. H; Mlsna. I. vv: Liebman. M; Mook. B; Grear.
J: Candelmo. AC: Chambers. RC: Gobler. CJ: Hunt. CW: King. AL; Price. NN;
Signorini. SR; Standoff. E; Stymies! C: Wahle. RA; Waller. JD; Rebuck. ND; Wang.
ZA; Capson, TL; Morrison, JR; Cooley, SR; Doney, SC. (2015). Ocean and Coastal
Acidification off New England and Nova Scotia. Oceanography 28: 182-197.
http://dx.doi.org/10.5670/oceanog.2015.41.
Glibert. PM; Allen. JI; Bouwman. AF; Brown. CW: Flynn. KJ; Lewitus. AJ; Madden. CJ.
(2010a). Modeling of HABs and eutrophication: Status, advances, challenges. J Mar Syst
83: 262-275. http://dx.doi.Org/10.1016/i.imarsvs.2010.05.004.
Glibert. PM: Bover. JN; Heil. CA; Madden. CJ: Sturgis. B; Wazniak. CS. (2010b). Blooms in
lagoons: Different from those of river-dominated estuaries. In MJ Kennish; HW Paerl
(Eds.), Coastal Lagoons: Critical Habitats of Environmental Change (pp. 91-113). Boca
Raton, FL: CRC Press. https://www.crcpress.eom/Coastal-Lagoons-Critical-Habitats-of-
Environmental-Change/Kenni sh-Paerl/97814200883 04.
Glibert. PM: Burford. MA. (2017). Globally changing nutrient loads and harmful algal blooms
recent advances, new paradigms, and continuing challenges. Oceanography 30: 58-69.
http://dx.doi.org/10.5670/oceanog.2017.110.
Glibert. PM; Hinkle. DC; Sturgis. B; Jesien. RV. (2014). Eutrophi cation of a Maryland/Virginia
coastal lagoon: A tipping point, ecosystem changes, and potential causes. Estuaries
Coasts 37: S128-S146. http://dx.doi.org/10.1007/s 12237-013-9630-3.
Glibert. PM; Wilkerson. FP; Dugdale. RC; Raven. JA; Dupont. CL; Leavitt. PR; Parker. AE;
Burkholder. JM; Kana. TM. (2016). Pluses and minuses of ammonium and nitrate uptake
and assimilation by phytoplankton and implications for productivity and community
composition, with emphasis on nitrogen-enriched conditions. Limnol Oceanogr 61: 165-
197. http://dx.doi.org/10.1002/lno.102Q3.
Glimour. CC. (2011). A review of the literature on the impact of sulfate on methylmercury in
sediments and soils. Edgewater, MD: Smithsonian Environmental Research Center.
Gobler. CJ; Burkholder. JM; Davis. TW; Harke. MJ; Johengen. T; Stow. CA; Van de Waal. DB.
(2016). The dual role of nitrogen supply in controlling the growth and toxicity of
cyanobacterial blooms [Review], Harmful Algae 54: 87-97.
http://dx.doi.Org/10.1016/i.hal.2016.01.010.
Gobler. CJ; Depasquale. EL; Griffith. AW; Baumann. H. (2014). Hypoxia and acidification have
additive and synergistic negative effects on the growth, survival, and metamorphosis of
early life stage bivalves. PLoS ONE 9: e83648.
http://dx.doi.org/10.1371/iournal.pone.0083648.
Gohre. V; Paszkowski. U. (2006). Contribution of the arbuscular mycorrhizal symbiosis to heavy
metal phytoremediation. Planta223: 1115-1122. http://dx.doi.org/10.1007/sQ0425-006-
0225-0.
June 2018
R-82
DRAFT: Do Not Cite or Quote

-------
Goldman. CR. (1988). Primary productivity, nutrients, and transparency during the early onset of
eutrophication in ultra-oligotrophic Lake Tahoe, California-Nevada. Limnology 33:
1321-1333. http://dx.doi.Org/10.4319/lo.1988.33.6.1321.
Goldman Mart one. R; Wasson. K. (2008). Impacts and interactions of multiple human
perturbations in a California salt marsh. Oecologia 158: 151-163.
http://dx.doi. org/10.1007/s00442-008-1129-4.
Gomot-De Vaufleury. A: Kerhoas. I. (2000). Effects of cadmium on the reproductive system of
the land snail Helix aspersa. Bull Environ Contam Toxicol 64: 434-442.
Goni-Urriza, M; Corsellis, Y; Lanceleur, L; Tessier, E; Gury, J; Monperrus, M; Guyoneaud, R.
(2015). Relationships between bacterial energetic metabolism, mercury methylation
potential, and hgcA/hgcB gene expression in Desulfovibrio dechloroacetivorans BerOcl.
Environ Sci Pollut Res Int 22: 13764-13771. http://dx.doi.org/10.1007/sll356-Q15-4273-
5.
Gonzalez. JA; Prado. FE; Piacentini. RD. (2014). Atmospheric dust accumulation on native and
non-native species: effects on gas exchange parameters. J Environ Qual 43: 801-808.
http://dx.doi.org/10.2134/ieq2013.08.03Q8.
Goodale, CL; Aber, JD. (2001). The long-term effects of land-use history on nitrogen cycling in
northern hardwood forests. Ecol Appl 11: 253-267. http://dx.doi.org/10.189Q/1051-
0761(200 n01ir0253:TLTEOL12.0.CQ:2.
Goodale. CL: Thomas. SA; Fredriksen. G: Elliott. EM: Flinn. KM: Butler. TJ; Walter. MT.
(2009). Unusual seasonal patterns and inferred processes of nitrogen retention in forested
headwaters of the Upper Susquehanna River. Biogeochemistry 93: 197-218.
http://dx.doi.org/10.1007/slQ533-009-9298-8.
Goodridge. BM; Melack. JM. (2012). Land use control of stream nitrate concentrations in
mountainous coastal California watersheds. J Geophys Res 117.
http://dx.doi.org/10.1029/2011JG0Q1833.
Gotelli. NJ: Ellison. AM. (2002). Nitrogen deposition and extinction risk in the northern pitcher
plant, Sarraceniapurpurea. Ecology 83: 2758-2765. http://dx.doi.org/10.1890/0Q12-
9658(2002)083r2758:NDAERn2.0.CQ;2.
Gotelli. NJ: Ellison. AM. (2006). Forecasting extinction risk with nonstationary matrix models.
Ecol Appl 16: 51-61. http://dx.doi.org/10.1890/04-0479.
Gough. L; Osenberg. CW: Gross. KL; Collins. SL. (2000). Fertilization effects on species
density and primary productivity in herbaceous plant communities. Oikos 89: 428-439.
http://dx.doi.Org/10.1034/i.1600-0706.2000.890302.x.
Goulder, LH; Kennedy, D. (1997). Valuing ecosystem services: Philosophical bases and
empirical methods. In GC Daily (Ed.), Nature's Services: Societal Dependence on
Natural Ecosystems (pp. 23-47). Washington, DC: Island Press.
Graham. AM: Aiken. GR; Gilmour. CC. (2012). Dissolved organic matter enhances microbial
mercury methylation under sulfidic conditions. Environ Sci Technol 46: 2715-2723.
http://dx.doi.org/10.1021/es203658f.
June 2018
R-83
DRAFT: Do Not Cite or Quote

-------
Graham. JL; Lot tin. KA; Meyer. MT; Ziegler. AC. (2010). Cyanotoxin mixtures and taste-and-
odor compounds in cyanobacterial blooms from the Midwestern United States. Environ
Sci Technol 44: 7361-7368. http://dx.doi.org/10.1021/eslQ08938.
Graham. SA; Mendelssohn. IA. (2010). Multiple levels of nitrogen applied to an oligohaline
marsh identify a plant community response sequence to eutrophication. Mar Ecol Prog
Ser417: 73-82. http://dx.doi.org/10.3354/meps08808.
Graham. SA: Mendelssohn. IA. (2016). Contrasting effects of nutrient enrichment on below-
ground biomass in coastal wetlands. J Ecol 104: 249-260. http://dx.doi.org/10. Ill 1/1365-
2745.12498.
Grantz. DA: Garner. JHB; Johnson. DW. (2003). Ecological effects of particulate matter.
Environ Int 29: 213-239. http://dx.doi.org/10.1016/SO160-4120(02^)00181-2.
Grantz. EM: Haggard. BE: Scott. JT. (2014). Stoichiometric imbalance in rates of nitrogen and
phosphorus retention, storage, and recycling can perpetuate nitrogen deficiency in highly-
productive reservoirs. Limnol Oceanogr 59: 2203-2216.
http://dx.doi.Org/10.4319/lo.2014.59.6.2203.
Gratao. PL: Polle. A: Lea. PJ: Azevedo. RA. (2005). Making the life of heavy metal-stressed
plants a little easier. Funct Plant Biol 32: 481-494. http://dx.doi.org/10.1071/FP05Q16.
Gray. DK; Arnott. SE. (2009). Recovery of acid damaged zooplankton communities:
Measurement, extent, and limiting factors. Environ Rev 17: 81-99.
http://dx.doi.Org/10.l 139/A09-006.
Gray. DK: Arnott. SE. (2011). Does dispersal limitation impact the recovery of zooplankton
communities damaged by a regional stressor? Ecol Appl 21: 1241-1256.
http://dx.doi.Org/10.1890/10-0364.l.
Gray. DK: Arnott. SE: Shead. JA: Perry. AM. (2012). The recovery of acid-damaged
zooplankton communities in Canadian Lakes: The relative importance of abiotic, biotic
and spatial variables. FreshwBiol 57: 741-758. http://dx.doi.org/10.Ill 1/i.1365-
2427.2012.02739.x.
Greaver. TL: Clark. CM: Compton. JE: Vallano. D: Talhelm. AF; Weaver. CP: Band. LE: Baron.
JS: Davidson. EA: Tague. CL: Felker-Quinn. E: Lynch. JA: Herrick. JD: Liu. L:
Goodale. CL: Novak. KJ: Haeuber. RA. (2016). Key ecological responses to nitrogen are
altered by climate change. Nat Clim Chang 6: 836-843.
http://dx.doi.org/10.1038/NCLIMATE3088.
Greaver. TL: Liu. L: Bobbink. R. (2011). Wetlands. In Assessment of nitrogen deposition effects
and empirical critical loads of nitrogen for ecoregions of the United States (General
Technical Report NRS-80) (pp. 193-208). Newtown Square, PA: U.S. Department of
Agriculture, Forest Service, Northern Research Station.
http://www.nrs.fs.fed.us/pubs/gtr/gtr-nrs-80chapters/17-greaver.pdf.
Greaver. TL: Sullivan. TJ: Herrick. JD: Barber. MC: Baron. JS: Cosby. BJ: Deerhake. ME:
Dennis. RL: Dubois. J. -JB: Goodale. CL: Herlihy. AT: Lawrence. GB: Liu. L: Lynch.
J A: Novak. KJ. (2012). Ecological effects of nitrogen and sulfur air pollution in the US:
What do we know? Front Ecol Environ 10: 365-372. http://dx.doi.Org/10.1890/l 10049.
June 2018
R-84
DRAFT: Do Not Cite or Quote

-------
Green. EK; Galatowitsch. SM. (2002). Effects of Phalaris arundinacea and nitrate-N addition on
the establishment of wetland plant communities. J Appl Ecol 39: 134-144.
Green. MB; Bailey. AS; Bailey. SW; Battles. JJ; Campbell JL; Driscoll, CT; Fahev, TJ; Lepine,
LC; Likens. GE; Ollinger, SV; Schaberg, PG. (2013). Decreased water flowing from a
forest amended with calcium silicate. Proc Natl Acad Sci USA 110: 5999-6003.
http://dx.doi.org/10.1073/pnas.13Q2445110.
Greene. RM; Lehrter. JC; Hagy. JD. III. (2009). Multiple regression models for hindcasting and
forecasting midsummer hypoxia in the Gulf of Mexico. Ecol Appl 19: 1161-1175.
http://dx.doi.Org/10.1890/08-0035.l.
Greening. H; Janicki. A. (2006). Toward reversal of eutrophic conditions in a subtropical
estuary: Water quality and seagrass response to nitrogen loading reductions in Tampa
Bay, Florida, USA. Environ Manage 38: 163-178. http://dx.doi.org/10.1007/sQ0267-005-
0079-4.
Greening. H; Janicki. A; Sherwood. ET; Pribble. R; Johansson. JOR. (2014). Ecosystem
responses to long-term nutrient management in an urban estuary: Tampa Bay, Florida,
USA. Estuar Coast Shelf Sci 151: A1-A16. http://dx.doi.Org/10.1016/i.ecss.2014.10.003.
Greening, HS; Cross, LM; Sherwood, ET. (2011). A multiscale approach to seagrass recovery in
Tampa Bay, Florida. Ecological Restoration 29: 82-93. http://dx.doi.org/10.3368/er.29.l-
2.82.
Greven, HC. (1992). Changes in the moss flora of the Netherlands. Biol Conserv 59: 133-137.
http://dx.doi.Org/l 0.1016/0006-3207(92)90572-5.
Griepentrog, M; Bode, S; Boeckx, P; Hagedorn, F; Heim, A; Schmidt, MWI. (2014). Nitrogen
deposition promotes the production of new fungal residues but retards the decomposition
of old residues in forest soil fractions. Global Change Biol 20: 327-340.
http://dx.doi.org/10. Ill 1/gcb. 12374.
Griepentrog, M; Eglinton, TI; Hagedorn, F; Schmidt, MWI; Wiesenberg, GLB. (2015).
Interactive effects of elevated C02 and nitrogen deposition on fatty acid molecular and
isotope composition of above- and belowground tree biomass and forest soil fractions.
Global Change Biol 21: 473-486. http://dx.doi.org/10. Ill 1/gcb. 12666.
Gries, C; Sanz, M, -J; Romagni, JG; Goldsmith, S; Kuhn, U; Kesselmeier, J; Nash, TH, III.
(1997). The uptake of gaseous sulphur dioxide by non-gelatinous lichens. New Phytol
135: 595-602. http://dx.doi.Org/10.1046/i.1469-8137.1997.00679.x.
Griffith, RS. (1991). Fraxinus americana. In: Fire Effects Information System. Available online
at http://www.fs.fed.us/database/feis/plants/tree/fraame/all.html (accessed March 16,
2016).
Grill, D. (1971). [Pufferkapazitat gesunder und rauchgeschadigter Fichtennadeln], J Plant Dis
Protect 78: 612-622.
Grime, JP. (1973). Competitive exclusion in herbaceous vegetation. Nature 242: 344-347.
http://dx.doi.org/10.1038/242344a0.
Grimm, NB; Fisher, SG. (1986). Nitrogen limitation in a Sonoran Desert stream. J North Am
Benthol Soc 5: 2-15. http://dx.doi.org/10.2307/1467743.
June 2018
R-85
DRAFT: Do Not Cite or Quote

-------
Grin an. E; Robinson. TMP. (2013). Resource availability and imbalance affect plant-mycorrhizal
interactions: A field test of three hypotheses. Ecology 94: 62-71.
http://dx.doi.ore/10.1890/12-03 85.1.
GrofYman. PM. (1994). Denitrification in freshwater wetlands. Current Topics Wetland
Biogeochem 1: 15-35.
GrofYman. PM: Driscoll. CT; Fahev. TJ; Hardy. JP; Fitzhugh, RD; Tiernev. GL. (2001). Effects
of mild winter freezing on soil nitrogen and carbon dynamics in a northern hardwood
forest. Biogeochemistry 56: 191-213.
Groffman, PM; Hardy, JP: Fashu-Kanu, S; Driscoll, CT; Cleavitt, NL; Fahey, TJ: Fisk, MC.
(2011). Snow depth, soil freezing and nitrogen cycling in a northern hardwood forest
landscape. Biogeochemistry 102: 223-238. http://dx.doi.org/10.1007/slQ533-010-9436-3.
Groffman, PM: Kareiva, P; Carter, S: Grimm, NB; Lawler, J: Mack, M; Matzek, V: Tallis, H.
(2014). Ch. 8: Ecosystems, biodiversity, and ecosystem service. In JM Melillo; TC
Richmond; GW Yohe (Eds.), Climate change impacts in the United States: The third
National Climate Assessment (pp. 195-219). Washington, DC: U.S. Global Change
Research Program. http://dx.doi.org/10.7930/J0TD9V7H.
Gronholm, T; Aalto, PP; Hiltunen, V: Rannik, U; Rinne, J: Laakso, L; Hyvonen, S; Vesala, T;
Kulmala, M. (2007). Measurements of aerosol particle dry deposition velocity using the
relaxed eddy accumulation technique. Tellus B Chem Phys Meteorol 59: 381-386.
http://dx.doi.org/10.1111/i. 1600-0889.2007.00268.X.
Gross, KL; Willig, MR; Gough, L; Inouye, R; Cox, SB. (2000). Patterns of species density and
productivity at different spatial scales in herbaceous plant communities. Oikos 89: 417-
427. http://dx.doi.Org/10.1034/i.1600-0706.2000.890301.x.
Gruber, N: Hauri, C: Lachkar, Z; Loher, D; Frolicher, TL; Plattner, GK. (2012). Rapid
progression of ocean acidification in the California Current System. Science 337: 220-
223. http://dx.doi.Org/10.l 126/science. 1216773.
Grulke, NE; Dobrowolski, W: Mingus, P; Fenn, ME. (2005). California black oak response to
nitrogen amendment at a high 03, nitrogen-saturated site. Environ Pollut 137: 536-545.
http://dx.doi.Org/10.1016/i.envpol.2005.01.039.
Grulke, NE: Minnich, RA; Paine, TP: Sevbold, SJ: Chavez, DJ; Fenn, ME: Riggan, PJ; Dunn, A.
(2008). Air pollution increases forest susceptibility to wildfires: A case study in the San
Bernardino Mountains in southern California. In A Bytnerowicz; MJ Arbaugh; AR
Riebau; C Anderson (Eds.), Wildland Fires and Air Pollution Developments in
Environmental Science (pp. 365-403). Amsterdam, The Netherlands: Elsevier Ltd.
http ://dx. doi .org/10.1016/S1474-8177(08)00017-X.
Gudasz, C: Bastviken, D; Steger, K; Premke, K; Sobek, S: Tranvik, LJ. (2010). Temperature-
controlled organic carbon mineralization in lake sediments. Nature 466: 478-481.
http://dx.doi.org/10.1038/nature09383.
Guimaraes, JRD; Mauro, JBN: Meili, M; Sundbom, M; Haglund, AL; Coelho-Souza, SA;
Hylander, LP. (2006). Simultaneous radioassays of bacterial production and mercury
methylation in the periphyton of a tropical and a temperate wetland. J Environ Manage
81: 95-100. http://dx.doi.Org/10.1016/i.ienvman.2005.09.023.
June 2018
R-86
DRAFT: Do Not Cite or Quote

-------
Gulati. RD; Demott WR. (1997). The role of food quality for zooplankton: Remarks on the
state-of-the-art, perspectives and priorities. FreshwBiol 38: 753-768.
http://dx.doi.Org/10.1046/i.1365-2427.1997.00275.x.
Gulmon. SL; Chu. CC. (1981). The effects of light and nitrogen on photosynthesis, leaf
characteristics, and dry-matter allocation in the chaparral shrub, diplacus-aurantiacus.
Oecologia 49: 207-212. http://dx.doi.org/10.1007/BF0Q349189.
Gundale. MJ; Deluca. TH; Nordin. A. (2011). Bryophytes attenuate anthropogenic nitrogen
inputs in boreal forests. Global Change Biol 17: 2743-2753.
http://dx.doi.org/10.1111/i. 1365-2486.2011,02407.x.
Gundale. MJ: From. F; Bach. LH; Nordin. A. (2014). Anthropogenic nitrogen deposition in
boreal forests has a minor impact on the global carbon cycle. Global Change Biol 20:
276-286. http://dx.doi.org/10.1111/gcb. 12422.
Gundersen, P; Emmett, BA; Kj0naas, OJ; Koopmans, CJ; Tietema, A. (1998). Impact of nitrogen
deposition on nitrogen cycling in forests: A synthesis of NITREX data. For Ecol Manage
101: 37-55. http://dx.doi.org/10.1016/S0378-l 127(97^)00124-2.
Gunn. J: Keller. W: Negusanti. J: Potvin. R; Beckett. P; Winterhalder. K. (1995). Ecosystem
recovery after emission reductions: Sudbury, Canada. Water Air Soil Pollut 85: 1783-
1788. http://dx.doi.org/10.1007/BF00477238.
Gunn. JM; McMurtry. MJ: Casselman. JM; Keller. W: Powell. MJ. (1988). Changes in the fish
community of a limed lake near Sudbury, Ontario: Effects of chemical neutralization
or reduced atmospheric deposition of acids? Water Air Soil Pollut 41: 113-136.
Gunn. JM: Mills. KH. (1998). The potential for restoration of acid-damaged lake trout lakes.
RestorEcol 6: 390-397. http://dx.doi.Org/10.1046/i.1526-100X.1998.06409.x.
Gunn. JM: Noakes. DLG. (1987). Latent effects of pulse exposure to aluminum and low pH on
size, ionic composition and feeding efficiency of lake trout (Salvelinus fontinalis)
alevins. Can J Fish Aquat Sci 44: 125-132. http://dx.doi.org/10.1139/f87-170.
Gtinthardtgoerg. MS: Schmutz. P; Matvssek. R; Bucher. JB. (1996). Leaf and stem structure of
poplar (Populus x euramericana) as influenced by 0-3, N02, their combination, and
different soil N supplies. Can J For Res 26: 649-657. http://dx.doi.org/10.1139/x26-075.
Gurevitch. J: Curtis. PS: Jones. MH. (2001). Meta-analysis in ecology. In Adv Ecol Res.
http://dx.doi.org/10.1016/S0065-2504(01B2013-5.
Guv. AL; Mischkolz. JM: Lamb. EG. (2013). Limited effects of simulated acidic deposition on
seedling survivorship and root morphology of endemic plant taxa of the Athabasca Sand
Dunes in well-watered greenhouse trials. Botany 91: 176-181.
http://dx.doi.Org/10.l 139/cib-2012-0162.
Gwiazda. RH; Broecker. WS. (1994). The separate and combined effects of temperature,
soilpC02, and organic acidity on silicate weathering in the soil environment: Formulation
of a model and results. Global Biogeochem Cycles 8: 141-155.
http://dx.doi.org/10.1029/94GB00491.
June 2018
R-87
DRAFT: Do Not Cite or Quote

-------
HABHRC Act. Harmful Algal Bloom and Hypoxia Research and Control Amendments Act:
Northern Gulf of Mexico hypoxia. 33 USC (2014).
http://uscode. house. gov/view.xhtml?req=granuleid:USC-prelim-title33-
section4004&num=0&edition=prelim
Hackney. CT; Avery. GB. (2015). Tidal wetland community response to varying levels of
flooding by saline water. Wetlands 35: 227-236. http://dx.doi.org/10.1007/sl3157-Q14-
0597-z.
Haddad. KD. (1989). Habitat trends and fisheries in Tampa and Sarasota Bays. In ED Estevez
(Ed.), Tampa and Sarasota bays: Issues, resources, status, and management (pp. 113-
128). Washington, DC: National Oceanic and Atmospheric Administration.
Hadlev. KR; Paterson. AM: Hall. RI; Smol. JP. (2013). Effects of multiple stressors on lakes in
south-central Ontario: 15 years of change in lakewater chemistry and sedimentary diatom
assemblages. Aquat Sci 75: 349-360. http://dx.doi.org/10.1007/s00027-012-028Q-5.
Hadlev. KR: Paterson. AM: Reid. RA; Rusak. JA; Somers. KM: Ingram. R; Smol. JP. (2015).
Altered pH and reduced calcium levels drive near extirpation of native crayfish,
Cambarus bartonii, in Algonquin Park, Ontario, Canada. Freshw Sci 34: 918-932.
http://dx.doi.org/10.1086/681910.
Hagan. PL: Waldrop. TA; Reillv. M; Shearman. TM. (2015). Impacts of repeated wildfire on
long-unburned plant communities of the southern Appalachian Mountains. International
Journal of Wildland Fire 24: 911-920. http://dx.doi.org/10.1071/WF14143.
Hagedorn, F; Kammer, A; Schmidt, MWI; Goodale, CL. (2012). Nitrogen addition alters
mineralization dynamics of 13C-depleted leaf and twig litter and reduces leaching of
older DOC from mineral soil. Global Change Biol 18: 1412-1427.
http://dx.doi.org/10.1111/i. 1365-2486.2011,02603.x.
Haines-Young, R; Potschin, M. (2011). Common international classification of ecosystem
services (CICES): 2011 update. Nottingham, UK: Centre for Environmental
Management, University of Nottingham.
http://unstats.un.org/unsd/envaccounting/seeaLES/egm/Issue8a.pdf.
Haines, TA; Baker, JP. (1986). Evidence of fish population responses to acidification in the
eastern United States. Water Air Soil Pollut 31: 605-629.
http://dx.doi.org/10.1007/BF0Q284216.
Haith, DA: Shoenaker, LL. (1987). Generalized watershed loading functions for stream flow
nutrients. Water Resour Bull 23: 471-478. http://dx.doi.org/10.Ill 1/i.1752-
1688.1987.tb00825.x.
Hall, BP: Aiken, GR; Krabbenhoft, DP: Marvin-Dipasquale, M; Swarzenski, CM. (2008).
Wetlands as principal zones of methylmercury production in southern Louisiana and the
Gulf of Mexico region. Environ Pollut 154: 124-134.
http://dx.doi.Org/10.1016/i.envpol.2007.12.017.
Hall, BP: Baron, LA: Somers, CM. (2009a). Mercury concentrations in surface water and
harvested waterfowl from the prairie pothole region of Saskatchewan. Environ Sci
Technol 43: 8759-8766. http://dx.doi.org/10.1021/es9Q24589.
June 2018
R-88
DRAFT: Do Not Cite or Quote

-------
Hall JL. (2002). Cellular mechanisms for heavy metal detoxification and tolerance. J Exp Bot
53: 1-11. http://dx.doi.Org/10.1093/iexbot/53.366.l.
Hall R.O. Jr; Baker. MA: Arp. CD: Koch. BJ. (2009b). Hydrologic control of nitrogen removal,
storage, and export in a mountain stream. Limnol Oceanogr 54: 2128-2142.
http://dx.doi.Org/10.4319/lo.2009.54.6.2128.
Hall. SJ: Sponseller. RA; Grimm. NB; Huber. D; Kave. JP; Clark. C: Collins. SL. (2011).
Ecosystem response to nutrient enrichment across an urban airshed in the Sonoran
Desert. Ecol Appl 21: 640-660. http://dx.doi.Org/10.1890/10-0758.l.
Hallingback, T. (1991). [Blue-green algae and cyanophilic lichens threatened by air pollution
and fertilization], Svensk Bot Tidskr 85: 87-104.
Hallingback. T; Kellner. O. (1992). Effects of simulated nitrogen rich and acid rain on the
nitrogen-fixing lichen Peltigera aphthosa (L.) Willd. New Phytol 120: 99-103.
http://dx.doi.Org/10.llll/i.1469-8137.1992.tb01062.x.
Halman. JM; Schaberg. PG: Hawlev. GJ: Eagar. C. (2008). Calcium addition at the Hubbard
Brook Experimental Forest increases sugar storage, antioxidant activity and cold
tolerance in native red spruce (Picea rubens). Tree Physiol 28: 855-862.
http://dx.doi.Org/10.1093/treephys/28.6.855.
Halman. JM: Schaberg. PG: Hawlev. GJ: Hansen. CF; Fahev. TJ. (2015). Differential impacts of
calcium and aluminum treatments on sugar maple and American beech growth dynamics.
Can J For Res 45: 52-59. http://dx.doi.Org/10.l 139/cifr-2014-0250.
Halman. JM: Schaberg. PG: Hawlev. GJ: Pardo. LH; Fahev. TJ. (2013). Calcium and aluminum
impacts on sugar maple physiology in a northern hardwood forest. Tree Physiol 33: 1242-
1251. http://dx.doi.org/10.1093/treephys/tpt099.
Hames. RS: Rosenberg. KV: Lowe. JD: Barker. SE: Dhondt. AA. (2002). Adverse effects of acid
rain on the distribution of the Wood Thrush Hylocichla mustelina in North America. Proc
Natl Acad Sci USA 99: 11235-11240. http://dx.doi.org/10.1073/pnas.172700199.
Han, H; Allan, JD. (2008). Estimation of nitrogen inputs to catchments: Comparison of methods
and consequences for riverine export prediction. Biogeochemistry 91: 177-199.
http://dx.doi.org/10.1007/slQ533-008-9279-3.
Hand, JL: Gebhart, KA: Schichtel, BA: Malm, WC. (2012a). Increasing trends in wintertime
particulate sulfate and nitrate ion concentrations in the Great Plains of the United States
(20002010). Atmos Environ 55: 107-110.
http://dx.doi.Org/10.1016/i.atmosenv.2012.03.050.
Hand, JL: Schichtel, BA: Malm, WC: Pitchford, ML. (2012b). Particulate sulfate ion
concentration and S02 emission trends in the United States from the early 1990s through
2010. Atmos Chem Phys 12: 10353-10365. http://dx.doi.org/10.5194/acp-12-lQ353-
2012.
Hand, JL: Schichtel, BA: Pitchford, M: Malm, WC: Frank, NH. (2012c). Seasonal composition
of remote and urban fine particulate matter in the United States. J Geophys Res Atmos
117. http://dx.doi.org/10.1029/2011JD017122.
June 2018
R-89
DRAFT: Do Not Cite or Quote

-------
Handa, IT; Aerts, R; Berendse, F; Berg. MP; Binder. A; Butenschoen, O: Chauvet E; Gessner,
MO; Jabiol. J; Makkonen. M; Mckie. BG; Malmqvist, B; Peeters. ETH. M; Scheu. S;
Schmid, B; van Ruijven, J; Vos, VCA; Haettenschwiler, S. (2014). Consequences of
biodiversity loss for litter decomposition across biomes. Nature 509: 218-+.
http://dx.doi.org/10.1038/naturel3247.
Handlev. L; Altsman. D; DeMav. R. (2007). Seagrass status and trends in the northern Gulf of
Mexico: 19402002 (pp. 1-267). (Scientific Investigations Report 2006-5287, US EPA
855-R-04-003). Reston, VA: Seagrass status and trends in the northern Gulf of Mexico:
19402002 U.S. DOI, U.S. Geological Survey, http://pubs.usgs.gov/sir/2006/5287/.
Hansen. DA; Edgerton. ES; Hartsell. BE; Jansen. JJ; Kandasamy. N; Hidv. GM; Blanchard. CL.
(2003). The southeastern aerorsol research and characterization study: Part I - overview. J
Air Waste Manag Assoc 53: 1460-1471.
http://dx.doi.org/10.1080/10473289.2003.10466318.
Hanson. A; Johnson. R; Wigand. C; Oczkowski. A; Davev. E; Markham. E. (2016). Responses
of Spartina alterniflora to Multiple Stressors: Changing Precipitation Patterns,
Accelerated Sea Level Rise, and Nutrient Enrichment. Estuaries Coasts 39: 1376-1385.
http://dx.doi.org/10.1007/sl2237-016-009Q-4.
Hanson. PJ; Amthor. JS; Wullschleger. SD; Wilson. KB; Grant. RF; Hartley. A; Hui. D; Hunt.
ER, Jr; Johnson, DW; Kimball, JS; King, AW; Luo, Y; McNulty, SG; Sun, G; Thornton,
PE; Wang, S; Williams, M; Baldocchi, DP; Cushman, RM. (2004). Oak forest carbon
and water simulations: Model intercomparisons and evaluations against independent data.
Ecol Monogr 74: 443-489. http://dx.doi.org/10.1890/03-4049.
Hanson, PJ; Garten, CT, Jr. (1992). Deposition of H15N03 vapour to white oak, red maple and
loblolly pine foliage: Experimental observations and a generalized model. New Phytol
122: 329-337. http://dx.doi.Org/10.llll/i.1469-8137.1992.tb04238.x.
Hanson, PJ; Lindberg, SE. (1991). Dry deposition of reactive nitrogen compounds: A review of
leaf, canopy and non-foliar measurements [Review], Atmos Environ 25: 1615-1634.
http://dx.doi.org/10.1016/0960-1686(91)90020-8.
Hanson, PJ; Rott, K; Taylor, GE, Jr; Gunderson, CA; Lindberg, SE; Ross-Todd, BM. (1989).
N02 deposition to elements representative of a forest landscape. Atmos Environ 23:
1783-1794. http://dx.doi.org/10.1016/0004-6981(89)90061-9.
Harding, LW, Jr; Batiuk, RA; Fisher, TR; Gallegos, CL; Malone, TC; Miller, WD; Mulholland,
MR; Paerl, HW; Perry, ES; Tango, P. (2014). Scientific bases for numerical chlorophyll
criteria in Chesapeake Bay. Estuaries Coasts 37: 134-148.
http://dx.doi.org/10.1007/sl2237-013-9656-6.
Harding, LW, Jr; Gallegos, CL; Perry, ES; Miller, WD; Adolf, JE; Mallonee, ME; Paerl, HW.
(2016a). Long-Term Trends of Nutrients and Phytoplankton in Chesapeake Bay.
Estuaries Coasts 39: 664-681. http://dx.doi.org/10.1007/sl2237-015-0Q23-7.
Harding, LW, Jr; Mallonee, ME; Perry, ES; Miller, WD; Adolf, JE; Gallegos, CL; Paerl, HW.
(2016b). Variable climatic conditions dominate recent phytoplankton dynamics in
Chesapeake Bay. Sci Rep 6: 23773. http://dx.doi.org/10.1038/srep23773.
June 2018
R-90
DRAFT: Do Not Cite or Quote

-------
Harpold. AA; Burns. DA; Walter. MT; Steenhuis. TS. (2013). Hydrogeomorphology explains
acidification-driven variation in aquatic biological communities in the Neversink Basin,
USA. Ecol Appl 23: 791-800. http://dx.doi.Org/10.1890/12-0603.l.
Harris. RC: Rudd. JWM; Amyot. M; Babiarz. CL; Beatv. K; Blanchfield. PJ: Bodalv. RA;
Branfireun. BA; Gilmour. CC: Gravdon. JA; Heves. A: Hintelmann. H; Hurley. JP;
Kelly. CA; Krabbenhoft. DP: Lindberg, SE; Mason. RP; Paterson. MJ; Podemski. CL:
Robinson, A, rt; Sandilands, K; Southworth, GR; Louis, V; Tate, MT. (2007). Whole-
ecosystem study shows rapid fish-mercury response to changes in mercury deposition.
Proc Natl Acad Sci USA 104: 16586-16591. http://dx.doi.org/10.1073/pnas.07041861Q4.
Harrison, JA: Maranger, RJ; Alexander, RB; Giblin, AE; Jacinthe, PA: Mavorga, E; Seitzinger,
SP; Sobota, DJ; Wollheim, WM. (2009). The regional and global significance of nitrogen
removal in lakes and reservoirs. Biogeochemistry 93: 143-157.
http://dx.doi.org/10.1007/slQ533-008-9272-x.
Hart, JA; Phlips, EJ; Badvlak, S: Dix, N; Petrinec, K; Mathews, AL; Green, W: Srifa, A. (2015).
Phytoplankton biomass and composition in a well-flushed, sub-tropical estuary: The
contrasting effects of hydrology, nutrient loads and allochthonous influences. Mar
Environ Res 112: 9-20. http://dx.doi.Org/10.1016/i.marenvres.2015.08.010.
Hartman, MP; Baron, JS; Ewing, HA; Weathers, KC. (2014). Combined global change effects
on ecosystem processes in nine US topographically complex areas. Biogeochemistry 119:
85-108. http://dx.doi.org/10.1007/slQ533-014-9950-9.
Hartman, MP; Baron, JS; Oiima, DS. (2007). Application of a coupled ecosystem-chemical
equilibrium model, DayCent-Chem, to stream and soil chemistry in a Rocky Mountain
watershed. Ecol Modell 200: 493-510.
http://dx.doi.Org/10.1016/i.ecolmodel.2006.09.001.
Harvey, HH; Lee, C. (1982). Historical fisheries changes related to surface water pH changes in
Canada. In RE Johnson (Ed.), Acid Rain/Fisheries: Proceedings of an International
Symposium on Acidic Precipitation and Fishery Impacts in Northeastern North America,
Cornell University, Ithaca, New York, August 2-5, 1981 (pp. 45-55). Bethesda, MD:
American Fisheries Society.
Hasler, AD. (1947). Eutrophication of lakes by domestic drainage. Ecology 28: 383-395.
http://dx.doi.org/10.2307/1931228.
Hassan, R; Scholes, R; Ash, N. (2005). Ecosystems and human well-being: Current state and
trends. In Ecosystems and Human Well-Being: Current State and Trends. Washington,
DC: Island Press.
Hasse, HE. (1913). The lichen flora of southern California. In The Lichen Flora of Southern
California. Department of Botany, Smithsonian Institution: Washington: G.P.O.
Hasselquist, NJ; Hogberg, P. (2014). Dosage and duration effects of nitrogen additions on
ectomycorrhizal sporocarp production and functioning: An example from two N-limited
boreal forests. Ecol Evol 4: 3015-3026. http://dx.doi.Org/10.1002/ece3.l 145.
June 2018
R-91
DRAFT: Do Not Cite or Quote

-------
Hasselquist NJ; Metcalfe. DB; Hogberg, P. (2012). Contrasting effects of low and high nitrogen
additions on soil C02 flux components and ectomycorrhizal fungal sporocarp production
in a boreal forest. Global Change Biol 18: 3596-3605.
http://dx.doi.org/10. Ill 1/gcb. 12001.
Hasselquist NJ: Metcalfe. DB: Marshall JD; Lucas. RW: Hogberg. P. (2016). Seasonality and
nitrogen supply modify carbon partitioning in understory vegetation of a boreal
coniferous forest. Ecology 97: 671-683. http://dx.doi.org/10.1890/15-0831.1.
Hastings Reserve. (2015). Lace lichen. Available online at
http://www.hastingsreserve.org/oakstory/LaceLichen2.html (accessed March 16, 2016).
Hattenrath. TK; Anderson. DM: Gobler. CJ. (2010). The influence of anthropogenic nitrogen
loading and meteorological conditions on the dynamics and toxicity of Alexandrium
fundyense blooms in a New York (USA) estuary. Harmful Algae 9: 402-412.
http://dx.doi.Org/10.1016/i.hal.2010.02.003.
Hauck. M; Wirth. V. (2010). Preference of lichens for shady habitats is correlated with
intolerance to high nitrogen levels. Lichenologist 42: 475-484.
http://dx.doi. org/10.1017/S0024282910000046.
Hauck, M; Zimmermann, J: Jacob, M; Dulamsuren, C; Bade, C; Ahrends, B; Leuschner, C.
(2012). Rapid recovery of stem increment in Norway spruce at reduced S02 levels in the
Harz Mountains, Germany. Environ Pollut 164: 132-141.
http://dx.doi.Org/10.1016/i.envpol.2012.01.026.
Hauri, C; Gruber, N; Plattner, GK; Alin, S; Feely, RA; Hales, B; Wheeler, PA. (2009). Ocean
acidification in the California Current System. Oceanography 22: 60-71.
http://dx.doi.org/10.5670/oceanog.2009.97.
Hautier, Y; Niklaus, PA: Hector, A. (2009). Competition for light causes plant biodiversity loss
after eutrophication. Science 324: 636-638. http://dx.doi.org/10.1126/science.l 169640.
Hautier, Y; Seabloom, EW: Borer, ET; Adler, PB; Harpole, WS: Hillebrand, H; Lind, EM:
Macdougall, AS: Stevens, CJ: Bakker, JD: Buckley, YM; Chu, C: Collins, SL; Daleo, P;
Damschen, EI; Davies, KF: Fay, PA: Firn, J: Gruner, DS: Jin, VL: Klein, JA: Knops,
JMH; La Pierre, KJ; Li, W; Mcculley, RL; Melbourne, BA; Moore, JL; O'Halloran, LR;
Prober, SM; Risch, AC; Sankaran, M; Schuetz, M; Hector, A. (2014). Eutrophi cation
weakens stabilizing effects of diversity in natural grasslands. Nature 508: 521-525.
http://dx.doi.org/10.1038/naturel3014.
Hauxwell, J; Cebrian, J; Valiela, I. (2003). Eelgrass Zostera marina loss in temperate estuaries:
Relationship to land-derived nitrogen loads and effect of light limitation imposed by
algae. Mar Ecol Prog Ser 247: 59-73. http://dx.doi.org/10.3354/meps247059.
Havas, M. (1986). Effects of acid deposition on aquatic ecosystems. In A Stern (Ed.), Air
Pollution, Volume VI: Supplement to Air Pollutants, Their Transformation, Transport,
and Effects (3rd ed. ed., pp. 351-389). New York, NY: Academic Press.
Havas, M; Woodfine, DG; Lutz, P; Yung, K; Maclsaac, HJ; Hutchinson, TC. (1995). Biological
recovery of two previously acidified, metal-contaminated lakes near Sudbury, Ontario,
Canada. Water Air Soil Pollut 85: 791-796. http://dx.doi.org/10.1007/BF0Q476926.
June 2018
R-92
DRAFT: Do Not Cite or Quote

-------
Havens. KE; Carlson. RE. (1998). Functional complementarity in plankton communities along a
gradient of acid stress. Environ Pollut 101: 427-436. http://dx.doi.org/10.1016/SQ269-
7491(98)00028-1.
Havn. M; Howarth. R; Marino. R; Ganiu. N: Berg. P; Foreman. KH; Giblin. AE; Mcglathery. K.
(2014). Exchange of Nitrogen and Phosphorus Between a Shallow Lagoon and Coastal
Waters. Estuaries Coasts 37: S63-S73. http://dx.doi.org/10.1007/sl2237-013-9699-8.
Haywood. JK. (1905). Injury to vegetation by smelter fumes. In Injury to Vegetation by Smelter
Fumes. Washington, DC: U.S. Department of Agriculture, Bureau of Chemistry.
He, C; Zhang, L; Demarchi, C; Croley, TE, II. (2014). Estimating point and non-point source
nutrient loads in the Saginaw Bay watersheds. J Great Lakes Res 40: 11-17.
http://dx.doi.Org/10.1016/i.iglr.2014.01.013.
He, Q: Silliman, BR. (2015). Biogeographic consequences of nutrient enrichment for plant-
herbivore interactions in coastal wetlands [Letter], Ecol Lett 18: 462-471.
http://dx.doi.org/10. Ill 1/ele. 12429.
He, S: Malfatti, SA; McFarland, JW: Anderson, FE; Pati, A: Huntemann, M; Tremblav, J: del
Rio, TG: Waldrop, MP: Windham-Myers, L; Tringe, SG. (2015). Patterns in wetland
microbial community composition and functional gene repertoire associated with
methane emissions. 6. http://dx.doi.Org/10.l 128/mBio.00066-15.
Heard, AM: Sickman, JO. (2016). Nitrogen assessment points: development and application to
high-elevation lakes in the Sierra Nevada, California. Ecosphere 7.
http://dx.doi.org/10.1002/ecs2.1586.
Heard, AM: Sickman, JO: Rose, NL; Bennett, DM: Lucero, DM: Melack, JM; Curtis, JH.
(2014). Twentieth-century atmospheric deposition and acidification trends in lakes of the
Sierra Nevada, California (USA). Environ Sci Technol 48: 10054-10061.
http://dx.doi.org/10.1021/es500934s.
Heath, J: Baron, JS. (2014). Climate, Not Atmospheric Deposition, Drives the Biogeochemical
Mass-Balance of a Mountain Watershed. Aquatic Geochemistry 20: 167-181.
http://dx.doi. org/10.1007/s 10498-013 -9199-2.
Hebertson, EG: Jenkins, MJ. (2008). Climate factors associated with historic spruce beetle
(Coleoptera: Curculionidae) outbreaks in Utah and Colorado. Environ Entomol 37: 281-
292. http://dx.doi.Org/10.1093/ee/37.2.281.
Hedin, LO: Armesto, JJ: Johnson, AH. (1995). Patterns of nutrient loss from unpolluted, old-
growth temperate forests: Evaluation of biogeochemical theory. Ecology 76: 493-509.
http://dx.doi.org/10.2307/1941208.
Hedwall, PO; Brunet, J: Rydin, H. (2017). Peatland plant communities under global change:
negative feedback loops counteract shifts in species composition. Ecology 98: 150-161.
http://dx.doi.org/10.1002/ecv.1627.
Hedwall, PO: Nordin, A: Strengbom, J: Brunet, J: Olsson, B. (2013). Does background nitrogen
deposition affect the response of boreal vegetation to fertilization? Oecologia 173: 615-
624. http://dx.doi.org/10.1007/s00442-013-2638-3.
June 2018
R-93
DRAFT: Do Not Cite or Quote

-------
Heinemever. A; Hartley. IP; Evans. SP; Carreira De La Fuente. JA; Ineson. P. (2007). Forest soil
C02 flux: Uncovering the contribution and environmental responses of ectomycorrhizas.
Global Change Biol 13: 1786-1797. http://dx.doi.org/10.111 1/j. 1365-2486.2007.01383 .x.
Heini. A: Puustinen. I; Tikka. M; Jokiniemi. A: Lepparanta. M; Arvola. L. (2014). Strong
dependence between phytoplankton and water chemistry in a large temperate lake:
Spatial and temporal perspective. Hydrobiologia 731: 139-150.
http://dx.doi.org/10.1007/slQ750-013-1777-l.
Heisler. J: Glibert. PM; Burkholder. JM; Anderson. DM: Cochlan. W: Dennison. WC: Dortch.
O; Gobler. CJ: Heil. CA; Humphries. E; Lewitus. A: Magnien. R; Marshall. HG: Sellner.
K; Stockwell. DA: Stoecker. DK; Suddleson. M. (2008). Eutrophication and harmful
algal blooms: A scientific consensus. Harmful Algae 8: 3-13.
http://dx.doi.Org/10.1016/i.hal.2008.08.006.
Heiss. EM: Fulweiler. RW. (2016). Coastal water column ammonium and nitrite oxidation are
decoupled in summer. Estuar Coast Shelf Sci 178: 110-119.
http://dx.doi.Org/10.1016/i.ecss.2016.06.002.
Hellal. J: Guedron. S: Huguet. L; Schaefer. J: Laperche. V: Joulian. C: Lanceleur. L; Burnol. A:
Ghestem. JP; Garrido. F; Battaglia-Brunet. F. (2015). Mercury mobilization and
speciation linked to bacterial iron oxide and sulfate reduction: A column study to mimic
reactive transfer in an anoxic aquifer. J Contam Hydrol 180: 56-68.
http://dx.doi.Org/10.1016/i.iconhyd.2015.08.001.
Helliwell. RC: Davies. JJL; Evans. CD: Jenkins. A: Coull. MC: Reynolds. B; Norris. D; Ferrier.
RC. (2007). Spatial and seasonal variations in nitrogen leaching and acidity across four
acid-impacted regions of the UK. Water Air Soil Pollut 185: 3-19.
http://dx.doi.org/10.1007/sll270-0Q7-9388-v.
Helmisaari. HS: Makkonen. K; Olsson. M; Viksna. A: Malkonen. E. (1999). Fine-root growth,
mortality and heavy metal concentrations in limed and fertilized Pinus silvestris (L.)
stands in the vicinity of a Cu-Ni smelter in SW Finland. Plant Soil 209: 193-200.
http://dx.doi.Org/10.1023/A:1004595531203.
Hember. RA; Kurz. WA: Coops. NC. (2017). Increasing net ecosystem biomass production of
Canada's boreal and temperate forests despite decline in dry climates. Global
Biogeochem Cycles 31: 134-158. http://dx.doi.org/10.1002/2016GB0Q5459.
Henry. EA; Dodge-Murphy. LJ; Bigham. GN: Klein. SM; Gilmour. CC. (1995). Total mercury
and methylmercury mass balance in an alkaline hypereutrophic urban lake (Onondaga
Lake, NY). Water Air Soil Pollut 80: 509517.
Henry. HAL: Hutchison. JS: Kim. M. inKu; McWhirter. BP. (2015). Context matters for
warming: Interannual variation in grass biomass responses to 7 years of warming and N
addition. Ecosystems 18: 103-114. http://dx.doi.org/10.1007/slQ021-014-9816-v.
Henrys. PA: Stevens. CJ: Smart. SM: Maskell. LC: Walker. KJ; Preston. CD: Crowe. A: Rowe.
EC; Gowing. DJ; Emmett. BA. (2011). Impacts of nitrogen deposition on vascular plants
in Britain: An analysis of two national observation networks. Biogeosciences 8: 3501-
3518. http://dx.doi.org/10.5194/bg-8-3501-2011.
June 2018
R-94
DRAFT: Do Not Cite or Quote

-------
Henze. DK; Seinfeld. JH; Shindell. DT. (2009). Inverse modeling and mapping US air quality
influences of inorganic PM2.5 precursor emissions using the adjoint of GEOS-Chem.
Atmos Chem Phys 9: 5877-5903.
Herlihy. AT; Kaufmann. PR; Church. MR; Wigington. PJ. Jr; Webb. JR; Sale. MJ. (1993). The
effects of acidic deposition on streams in the Appalachian Mountain and Piedmont
Region of the mid-Atlantic United States. Water Resour Res 29: 2687-2703.
http://dx.d0i.0rg/l 0.1029/93WRO1072.
Hernandez. PL; Vallano. DM; Zavaleta. ES; Tzankova. Z; Pasari. JR; Weiss. S; Selmants. PC;
Morozumi. C. (2016). Nitrogen pollution is linked to US listed species declines.
Bioscience 66: 213-222. http://dx.doi.org/10.1093/biosci/biw003.
Hernandez. L; Probst. A; Probst. JL; Ulrich. E. (2003). Heavy metal distribution in some French
forest soils: Evidence for atmospheric contamination. Sci Total Environ 312: 195-219.
http://dx.doi.org/10.1016/S0048-9697(03X)0223-7.
Herr. JW; Chen. CW. (2012). WARMF: MODEL USE, CALIBRATION, AND VALIDATION.
American Society of Agricultural and Biological Engineers Transactions 55: 1385-1394.
Hesse. CN; Mueller. RC; Vuvisich. M; Gallegos-Graves. LV; Gleasner. CD; Zak. PR; Kuske.
CR. (2015). Forest floor community metatranscriptomes identify fungal and bacterial
responses to N deposition in two maple forests. FMICB 6: 337.
http://dx.doi.org/10.3389/fmicb.2015.00337.
Hessen. P. (2013). Inorganic Nitrogen Peposition and Its Impacts on N:P-Ratios and Lake
Productivity. Water 5: 327-341. http://dx.doi.org/10.3390/w502Q327.
Hessing-Lewis. ML; Hacker. SP. (2013). Upwelling-influence, macroalgal blooms, and seagrass
production; temporal trends from latitudinal and local scales in northeast Pacific
estuaries. Limnol Oceanogr 58: 1103-1112. http://dx.doi.Org/10.4319/lo.2013.58.3.l 103.
Hessing-Lewis. ML; Hacker. SP; Menge. BA; Mcconville. S. eaoh; Henderson. J. (2015). Are
large macroalgal blooms necessarily bad? nutrient impacts on seagrass in upwelling-
influenced estuaries. Ecol Appl 25: 1330-1347. http://dx.doi.Org/10.1890/14-0548.l.
Hessing-Lewis. ML; Hacker. SP; Menge. BA; Rumrill. SS. (2011). Context-Pependent
Eelgrass-Macroalgae Interactions Along an Estuarine Gradient in the Pacific Northwest,
USA. Estuaries Coasts 34: 1169-1181. http://dx.doi.org/10.10Q7/sl2237-011-9412-8.
Hesthagen, T; Fiske, P; Saksgard, R. (2016). Recovery of young brown trout (Salmo trutta) in
acidified streams: What are the critical values for acid-neutralizing capacity? Atmos
Environ 146: 236-244. http://dx.doi.Org/10.1016/i.atmosenv.2016.07.010.
Hesthagen. T; Fiske. P; Skielkvale. BL. (2008). Critical limits for acid neutralizing capacity of
brown trout (Salmo trutta) in Norwegian lakes differing in organic carbon concentrations.
Aquatic Ecology 42: 307-316. http://dx.doi.org/10.1007/slQ452-008-9191-x.
Hesthagen. T; Larsen. BM; Fiske. P. (2011). Liming restores Atlantic salmon (Salmo salar)
populations in acidified Norwegian rivers. Can J Fish Aquat Sci 68: 224-231.
http://dx.doi.Org/10.l 139/F10-133.
June 2018
R-95
DRAFT: Do Not Cite or Quote

-------
Hettelingh, JP; Schiltze. G; de Vries. W; van der Gon. HP; Ilyin. I; Reinds. GJ; Slootweg, J;
Travnikov. O. (2015). Critical loads of cadmium, lead and mercury and their exceedances
in Europe. In W de Vries; JP Hettelingh; M Posch (Eds.), Critical loads and dynamic risk
assessments: Nitrogen, acidity and metals in terrestrial and aquatic ecosystems (pp. 523-
546). Bilthoven, The Netherlands: Springer Netherlands. http://dx.doi.org/10.10Q7/978-
94-017-9508-1 21.
Hettinger, A; Sanford, E; Hill, TM; Russell, AD; Sato, KNS; Hoey, J; Forsch, M; Page, HN;
Gavlord, B. (2012). Persistent carry-over effects of planktonic exposure to ocean
acidification in the Olympia oyster. Ecology 93: 2758-2768.
http://dx.doi.Org/10.1890/12-0567.l.
Heuschele, J; Candolin, U. (2010). Reversed parasite-mediated selection in sticklebacks from
eutrophied habitats. BehavEcol Sociobiol 64: 1229-1237.
http://dx.doi.org/10.1007/s00265-010-Q937-9.
Heuschele, J: Salminen, T; Candolin, U. (2012). Habitat change influences mate search
behaviour in three-spined sticklebacks. Anim Behav 83: 1505-1510.
http://dx.doi.Org/10.1016/i.anbehav.2012.03.027.
HEW (U.S. Department of Health, Education and Welfare). (1969). Air quality criteria for sulfur
oxides. Washington, DC: National Air Pollution Control Administration.
Hibben, CR; Silverborg, SB. (1978). Severity and causes of ash dieback. 4: 274-279.
Hickev, BM; Banas, NS. (2003). Oceanography of the U.S. Pacific Northwest Coastal Ocean and
estuaries with application to coastal ecology. Estuaries 26: 10101031.
http://dx.doi.org/10.1007/BF0280336Q.
Hicks, BB; Baldocchi, DP: Meyers, TP: Hosker, RP, Jr; Matt, DR. (1987). A preliminary
multiple resistance routine for deriving dry deposition velocities from measured
quantities. Water Air Soil Pollut 36: 311-330. http://dx.doi.org/10.1007/BF0Q229675.
Higgins, CB; Stephenson, K; Brown, BL. (2011). Nutrient bioassimilation capacity of
aquacultured oysters: Quantification of an ecosystem service. J Environ Qual 40: 271-
277. http://dx.doi.org/10.2134/iea2010.02Q3.
Hill, BH; Jicha, TM: Lehto, LL; Elonen, CM: Sebestven, SD; Kolka, RK. (2016). Comparisons
of soil nitrogen mass balances for an ombrotrophic bog and a minerotrophic fen in
northern Minnesota. Sci Total Environ 550: 880-892.
http://dx.doi.Org/10.1016/i.scitotenv.2016.01.178.
Hill, WR; Knight, AW. (1988). Nutrient and light limitation of algae in 2 northern California
streams. JPhycol 24: 125-132. http://dx.doi.Org/10.llll/i.1529-8817.1988.tb04225.x.
Hillerislambers, J; Harpole, WS; Schnitzer, S; Tilman, D; Reich, PB. (2009). C02, nitrogen, and
diversity differentially affect seed production of prairie plants. Ecology 90: 1810-1820.
http://dx.doi.Org/10.1890/07-1351.l.
Hines, DE: Lisa, JA; Song, B; Tobias, CR: Borrett, SR. (2012). A network model shows the
importance of coupled processes in the microbial N cycle in the Cape Fear River Estuary.
Estuar Coast Shelf Sci 106: 45-57. http://dx.doi.Org/10.1016/i.ecss.2012.04.018.
June 2018
R-96
DRAFT: Do Not Cite or Quote

-------
Hobbelen. PHF; Koolhaas. JE; van Gestel. CAM. (2006). Bioaccumulation of heavy metals in
the earthworms Lumbricus rubellus and Aporrectodea caliginosa in relation to total and
available metal concentrations in field soils. Environ Pollut 144: 639-646.
http://dx.doi.Org/10.1016/i.envpol.2006.01.019.
Hobbie, SE. (2008). Nitrogen effects on decomposition: A five-year experiment in eight
temperate sites. Ecology 89: 2633-2644. http://dx.doi.org/10.1890/07-l 119.1.
Hobbie. SE: Eddy. WC: Buvarski. CR; Adair. EC: Ogdahl. ML: Weisenhorn. P. (2012).
Response of decomposing litter and its microbial community to multiple forms of
nitrogen enrichment. Ecol Monogr 82: 389-405. http://dx.doi.Org/10.1890/l 1-1600.1.
Hobbs. R; Huenneke. LF. (1996). Disturbance, diversity, and invasion: Implications for
conservation. In FB Samaon; FL Knopf (Eds.), Ecosystem Management (pp. 164-180).
New York, NY: Springer New York, http://dx.doi.org/10.1007/978-l-4612-4018-l 16.
Hobbs, WO: Lafrancois, BM; Stottlemyer, R; Toczydlowski, D; Engstrom, PR; Edlund, MB;
Almendinger. JE: Strock. KE; Vandermeulen. D; Elias. JE: Saros. JE. (2016). Nitrogen
deposition to lakes in national parks of the western Great Lakes region: Isotopic
signatures, watershed retention, and algal shifts. Global Biogeochem Cycles 30: 514-533.
http://dx.doi.org/10.1002/2015GB0Q5228.
Hobbs. WO: Telford. RJ: Birks. HJ: Saros. JE: Hazewinkel. RR: Perren. BB: Saulnier-Talbot. E:
Wolfe. AP. (2010). Quantifying recent ecological changes in remote lakes of North
America and Greenland using sediment diatom assemblages. PLoS ONE 5: el0026.
http://dx.doi.org/10.1371/iournal.pone.0010Q26.
Hoellein. TJ: Zarnoch. CB: Grizzle. RE. (2015). Eastern oyster (Crassostrea virginica) filtration,
biodeposition, and sediment nitrogen cycling at two oyster reefs with contrasting water
quality in Great Bay Estuary (New Hampshire, USA). Biogeochemistry 122: 113-129.
http://dx.doi.org/10.1007/slQ533-014-0034-7.
Hogan. EJ: Minnullina. G: Sheppard. LJ: Leith. ID: Crittenden. PP. (2010a). Response of
phosphomonoesterase activity in the lichen Cladonia portentosa to nitrogen and
phosphorus enrichment in a field manipulation experiment. New Phytol 186: 926-933.
http://dx.doi.Org/10.llll/i.1469-8137.2010.03221.x.
Hogan. EJ: Minnullina. G: Smith. RI: Crittenden. PP. (2010b). Effects of nitrogen enrichment on
phosphatase activity and nitrogen:phosphorus relationships in Cladonia portentosa. New
Phytol 186: 911-925. http://dx.doi.org/10.111 l/i.l469-8137.2010.03222.x.
Hogberg, MN: Briones, MJI: Keel, SG: Metcalfe, PB: Campbell, C: Midwood, AJ: Thornton, B:
Hurry, V: Linder, S: Nasholm, T; Hogberg, P. (2010). Quantification of effects of season
and nitrogen supply on tree below-ground carbon transfer to ectomycorrhizal fungi and
other soil organisms in a boreal pine forest. New Phytol 187: 485-493.
http://dx.doi.org/10.111 l/i.l469-8137.2010.03274.x.
Hogberg, MN: Hogbom, L: Kleia, P, anB. (2013). Soil microbial community indices as
predictors of soil solution chemistry and N leaching in Picea abies (L.) Karst. forests in S.
Sweden. Plant Soil 372: 507-522. http://dx.doi.org/10.1007/sl 1104-013-1742-9.
June 2018
R-97
DRAFT: Do Not Cite or Quote

-------
Hogberg, P. (2012). What is the quantitative relation between nitrogen deposition and forest
carbon sequestration? Global Change Biol 18: 1-2. http://dx.doi.org/10.Ill 1/i.1365-
2486.2011.02553.x.
Hogberg, P; Fan. H; Quist M; Binklevs, D; Olof Tamm, C. (2006). Tree growth and soil
acidification in response to 30 years of experimental nitrogen loading on boreal forest.
Global Change Biol 12: 489-499. http://dx.doi.org/10.1111/i. 1365-2486.2006.01102.x.
Hogberg, P; Johannisson, C: Yarwood, S: Callesen, I; Nasholm, T; Myrold, DP: Hogberg, MN.
(2011). Recovery of ectomycorrhiza after 'nitrogen saturation' of a conifer forest. New
Phytol 189: 515-525. http://dx.doi.org/10.111 l/i.l469-8137.2010.03485.x.
Hoggarth. CGJ: Hall BP: Mitchell CP J. (2015). Mercury methylation in high and low-sulphate
impacted wetland ponds within the prairie pothole region of North America. Environ
Pollut 205: 269-277. http://dx.doi.Org/10.1016/i.envpol.2015.05.046.
Hogsden, KL; Xenopoulos, MA; Rusak, JA. (2009). Asymmetrical food web responses in
trophic-level richness, biomass, and function following lake acidification. Aquatic
Ecology 43: 591-606. http://dx.doi.org/10.1007/slQ452-008-9169-8.
Holland. EA; Dentener. FJ; Braswell BH; Sulzman. JM. (1999). Contemporary and pre-
industrial global reactive nitrogen budgets. Biogeochemistry 46: 7-43.
http://dx.doi.Org/10.1023/A:1006148011944.
Holland. S. ..Mansur. E.... M..uller. N.... Y..ates. A... (2015). Environmental benefits from
driving electric vehicles? 1. http://dx.doi.org/10.3386/w21291.
Hollweg. TA; Gilmour. CC: Mason. RP. (2009). Methylmercury production in sediments of
Chesapeake Bay and the mid-Atlantic continental margin. Mar Chem 114: 86-101.
http://dx.doi. org/10.1016/i .marchem .2009.04.004.
Hollweg. TA: Gilmour. CC: Mason. RP. (2010). Mercury and methylmercury cycling in
sediments of the mid-Atlantic continental shelf and slope. Limnol Oceanogr 55: 2703-
2722. http://dx.doi.Org/10.4319/lo.2010.55.6.2703.
Holmer, M; Thorsen, SW; Carlsson, MS: Kjerulf, PJ. (2015). Pelagic and benthic nutrient
regeneration processes in mussel cultures (Mytilus edulis) in a eutrophic coastal area
(Skive Fjord, Penmark). Estuaries Coasts 38: 1629-1641.
http://dx.doi.org/10.1007/sl2237-014-9864-8.
Holmes, JA; Franklin, EC: Gould, RA. (1915). Selby report: Report of the Selby Smelter
Commission. In Report on the Selby Smelter Commission. Washington, PC: Pepartment
of the Interior, Bureau of Mines.
Holmgren, K; Pegerman, E; Petersson, E; Bergquist, B. (2016). Long term trends of fish after
liming of Swedish streams and lakes. Atmos Environ 146: 245-251.
http://dx.doi.Org/10.1016/i.atmosenv.2016.08.033.
Holt, C: Yan, NP. (2003). Recovery of crustacean zooplankton communities from acidification
in Killarney Park, Ontario, 1971-2000: pH 6 as a recovery goal. R Swedish Acad Sci 32:
203-207. http://dx.doi.org/10.1639/0044-7447(2003)032r0203:ROCZCF12.0.CQ:2.
June 2018
R-98
DRAFT: Do Not Cite or Quote

-------
Holt CA; Yan. ND; Somers. KM. (2003). pH 6 as the threshold to use in critical load modeling
for zooplankton community change with acidification in lakes of south-central Ontario:
Accounting for morphometry and geography. Can J Fish Aquat Sci 60: 151-158.
http://dx.doi.Org/10.l 139/F03-008.
Holtgrieve, GW: Schindler. DE; Hobbs. WO: Leavitt PR: Ward. EJ; Bunting. L; Chen. G:
Finney. BP: Gregory-Eaves. I; Holmgren. S: Lisac. MJ: Lisi. PJ: Nvdick. K: Rogers. LA:
Saros, JE: Selbie, DT: Shapley, MP: Walsh, PB: Wolfe, AP. (2011). A coherent
signature of anthropogenic nitrogen deposition to remote watersheds of the Northern
Hemisphere. Science 334: 1545-1548. http://dx.doi.Org/10.l 126/science. 1212267.
Holtze, KE: Hutchinson, NJ. (1989). Lethality of low pH and A1 to early life stages of six fish
species inhabiting precambrian shield waters in Ontario. Can J Fish Aquat Sci 46: 1188-
1202. http://dx.doi.org/10.1139/f89-155.
Holzmueller, EJ: Jose, S: Jenkins, MA. (2007). Influence of calcium, potassium, and magnesium
on Cornus florida L. density and resistance to dogwood anthracnose. Plant Soil 290: 189-
199. http://dx.doi.org/10.1007/s 11104-006-9151 -v.
Homyak, PM; Sickman, JO: Miller, AE; Melack, JM; Meixner, T; Schimel, JP. (2014).
Assessing nitrogen-saturation in a seasonally dry Chaparral watershed: Limitations of
traditional indicators of N-saturation. Ecosystems 17: 1286-1305.
http://dx.doi.org/10.1007/slQ021-014-9792-2.
Hong, B: Swanev, DP: Howarth, RW. (2011). A toolbox for calculating net anthropogenic
nitrogen inputs (NANI). Environ Modell Softw 26: 623-633.
http://dx.doi.Org/10.1016/i.envsoft.2010.l 1.012.
Hong, B: Swanev, DP: Howarth, RW. (2013). Estimating Net Anthropogenic Nitrogen Inputs to
US Watersheds: Comparison of Methodologies. Environ Sci Technol 47: 5199-5207.
http://dx.doi.org/10.1021/es303437c.
Honour, SL: Bell, JN: Ashenden, TW: Cape, JN: Power, SA. (2009). Responses of herbaceous
plants to urban air pollution: effects on growth, phenology and leaf surface
characteristics. Environ Pollut 157: 1279-1286.
http://dx.doi.Org/10.1016/i.envpol.2008.l 1.049.
Hooda, V. (2007). Phytoremediation of toxic metals from soil and waste water. J Environ Biol
28: 367-376.
Hooper, DU: Adair, EC: Cardinale, BJ: Byrnes, JEK: Hungate, BA: Matulich, KL: Gonzalez, A:
Duffy, JE: Gamfeldt, L: O'Connor, MI. (2012). A global synthesis reveals biodiversity
loss as a major driver of ecosystem change. Nature 486: 105-U129.
http://dx.doi.org/10.1038/naturel 1118.
Hops, AB; McMahon, G. (2009). Spatial analysis of instream nitrogen loads and factors
controlling nitrogen delivery to streams in the southeastern United States using spatially
referenced regression on watershed attributes (SPARROW) and regional classification
frameworks. Hydrolog Process 23: 2275-2294. http://dx.doi.org/10.1002/hyp.7323.
June 2018
R-99
DRAFT: Do Not Cite or Quote

-------
Horowitz. LW; Walters. S; Mauzerall. PL; Emmons. LK; Rasch. PJ; Granier. C; Tie. X;
Lamarque. J. -F; Schultz. MG; Tyndall. GS; Orlando. JJ; Brasseur. GP. (2003). A global
simulation of tropospheric ozone and related tracers: Description and evaluation of
MOZART, version 2. J Geophys Res 108. http://dx.doi.org/10.1029/2002JD0Q2853.
Horslev. SB: Bailey. SW: Ristau. TE; Long. RP; Hallett. RA. (2008). Linking environmental
gradients, species composition, and vegetation indicators of sugar maple health in the
northeastern United States. Can J For Res 38: 1761-1774. http://dx.doi.org/10.1139/XQ8-
023.
Horswill. P; O'Sullivan. O; Phoenix. GK; Lee. JA; Leake. J. R. (2008). Base cation depletion,
eutrophication and acidification of species-rich grasslands in response to long-term
simulated nitrogen deposition. Environ Pollut 155: 336-349.
http://dx.doi.Org/10.1016/i.envpol.2007.l 1.006.
Horz. HP: Barbrook. A: Field. CB; Bohannan. BJM. (2004). Ammonia-oxidizing bacteria
respond to multifactorial global change. Proc Natl Acad SciUSAlOl: 15136-15141.
http://dx.doi.org/10.1073/pnas.04Q6616101.
Howarth. R; Chan. F; Conlev. DJ: Gamier. J: Donev. SC: Marino. R; Billen. G. (2011). Coupled
biogeochemical cycles: Eutrophi cation and hypoxia in temperate estuaries and coastal
marine ecosystems. Front Ecol Environ 9: 18-26. http://dx.doi.org/10.1890/1000Q8.
Howarth. R; Swanev. D; Billen. G: Gamier. J: Hong. B; Humborg. C: Johnes. P; Morth. C. -M;
Marino. R. (2012). Nitrogen fluxes from the landscape are controlled by net
anthropogenic nitrogen inputs and by climate. Front Ecol Environ 10: 37-43.
http://dx.doi.org/10.1890/100178.
Howarth. RW. (1988). Nutrient limitation of net primary production in marine ecosystems. Annu
Rev Ecol Evol Systemat 19: 89-110.
Howarth, RW. (2008a). Coastal nitrogen pollution: A review of sources and trends globally and
regionally. Harmful Algae 8: 14-20. http://dx.doi.Org/10.1016/i.hal.2008.08.015.
Howarth. RW. (2007). Atmospheric deposition and nitrogen pollution in coastal marine
ecosystems. In GR Visgilio; DM Whitelaw (Eds.), Acid in the Environment (pp. 97-116).
New York, NY: Springer, http://dx.doi.org/10.1007/978-0-387-37562-5 6.
Howarth. RW. (2008b). Estimating atmospheric nitrogen deposition in the Northeastern United
States: Relevance to Narragansett Bay. In A Desbonnet; BA Cost-Pierce (Eds.), Science
for ecosystem-based management: Narragansett Bay in the 21st century (pp. 47-65). New
York: Springer New York, http://dx.doi.org/10.1007/978-0-387-35299-2 3.
Howarth. RW: Billen. G: Swanev. D; Townsend. A: Jaworski. N: Laitha. K; Downing. JA:
Elmgren. R; Caraco. N: Jordan. T; Ferendse. F; Frenev. J: Kudevarov. V: Murdoch. P;
Zhao-Liang. Z. (1996a). Regional nitrogen budgets and riverine N & P fluxes for the
drainages to the North Atlantic Ocean: Natural and human influences. Biogeochemistry
35: 75-139. http://dx.doi.org/10.1007/BF02179825.
Howarth. RW: Bover. EW: Pabich. WJ: Galloway. JN. (2002). Nitrogen use in the United States
from 1961-2000 and potential future trends. Ambio 31: 88-96.
http://dx.doi.Org/10.1579/0044-7447-31.2.88.
June 2018
R-100
DRAFT: Do Not Cite or Quote

-------
Howarth. RW: Havn. M; Marino. RM; Ganiu. N; Foreman. K; Mcglathery. K; Giblin. AE; Berg.
P; Walker. JD. (2014). Metabolism of a nitrogen-enriched coastal marine lagoon during
the summertime. Biogeochemistry 118: 1-20. http://dx.doi.org/10.1007/slQ533-013-
9901-x.
Howarth. RW: Marino. R. (2006). Nitrogen as the limiting nutrient for eutrophication in coastal
marine ecosystems: Evolving views over three decades. Limnology 51: 364-376.
http://dx.doi.Org/10.4319/lo.2006.51.l part 2.0364.
Howarth. RW: Schneider. R; Swanev. D. (1996b). Metabolism and organic carbon fluxes in the
tidal freshwater Hudson River. Estuaries 19: 848-865. http://dx.doi.org/10.2307/1352302.
Howarth. RW: Swanev. DP: Bover. EW: Marino. R; Jaworski. N: Goodale. C. (2006). The
influence of climate on average nitrogen export from large watersheds in the
Northeastern United States. Biogeochemistry 79: 163-186.
http://dx.doi.org/10.1007/sl0533-006-901Q-l.
Howarth. RW: Swanev. DP: Butler. TJ; Marino. R. (2000). Climatic control on eutrophi cation of
the Hudson River estuary. Ecosystems 3: 210-215.
Hovt. DV. (1978). A model for the calculation of solar global insolation. Solar Energy 21: 27-35.
http ://dx. doi .org/10.1016/003 8-092X(78)90113-5.
Hrabik. TR: Watras. CJ. (2002). Recent declines in mercury concentration in a freshwater
fishery: Isolating the effects of de-acidification and decreased atmospheric mercury
deposition in Little Rock Lake. Sci Total Environ 297: 229-237.
http ://dx. doi .org/10.1016/S0048-9697(02)00138-9.
Hruska. J: Kram. P: McDowell. WH: Oulehle. F. (2009). Increased dissolved organic carbon
(DOC) in central European streams is driven by reductions in ionic strength rather than
climate change or decreasing acidity. Environ Sci Technol 43: 4320-4326.
http://dx.doi.org/10.1021/es803645w.
Hsu-Kim. H: Kucharzyk. KH: Zhang. T: Deshusses. MA. (2013). Mechanisms regulating
mercury bioavailability for methylating microorganisms in the aquatic environment: A
critical review [Review], Environ Sci Technol 47: 2441-2456.
http://dx.doi.org/10.1021/es304370g.
Hu. X: Pollack. JB: Mccutcheon. MR: Montagna. PA: Ouvang. Z. (2015a). Long-term alkalinity
decrease and acidification of estuaries in northwestern Gulf of Mexico. Environ Sci
Technol 49: 3401-3409. http://dx.doi.org/10.1021/es505945p.
Hu. Y. -L: Jung. K: Zeng. D. -H: Chang. SX. (2013). Nitrogen-and sulfur-deposition-altered soil
microbial community functions and enzyme activities in a boreal mixedwood forest in
western Canada. Can J For Res 43: 777-784. http://dx.doi.org/10.1139/cifr-2013-0049.
Hu. Y; Bellaloui. N: Tigabu. M: Wang. J: Diao. J: Wang. K. e: Yang. R. ui: Sun. G. (2015b).
Gaseous N02 effects on stomatal behavior, photosynthesis and respiration of hybrid
poplar leaves. Acta Physiologiae Plantarum 37: 1-8. http://dx.doi.org/10.10Q7/sll738-
014-1749-8.
June 2018
R-101
DRAFT: Do Not Cite or Quote

-------
Huang. G; Li. Y; Su. YG. (2015). Divergent responses of soil microbial communities to water
and nitrogen addition in a temperate desert. Geoderma 251-252: 55-64.
http://dx.doi.ore/10.1016/i.geoderma.2015.03.018.
Huang. L; Zhu. W: Ren. H; Chen. H; Wang. J. (2012). Impact of atmospheric nitrogen
deposition on soil properties and herb-layer diversity in remnant forests along an urban-
rural gradient in Guangzhou, southern China. Plant Ecol 213: 1187-1202.
http://dx.doi.org/10.1007/sll258-012-008Q-v.
Huang. S: Wang. X: Ma. G: He. G: Wu. J. (2017). The effects of aluminum stress on archaeal
diversity in red soils. Fresen Environ Bull 26: 1902-1910.
Hubbard. KA. Jr; Lautz. LK; Mitchell. MJ; Mayer. B; Hotchkiss. ER. (2010). Evaluating nitrate
uptake in a Rocky Mountain stream using labelled N-15 and ambient nitrate chemistry.
Hydrolog Process 24: 3322-3336. http://dx.doi.org/10.1002/hyp.7764.
Hudman, RC; Moore, NE; Mebust, AK; Martin, RV; Russell, AR; Valin, LC; Cohen, RC.
(2012).	Steps towards a mechanistic model of global soil nitric oxide emissions:
Implementation and space based-constraints. Atmos Chem Phys 12: 7779-7795.
http://dx.doi. org/10.5194/acp-12-7779-2012.
Huebert, DB; L'Hirondelle, SJ; Addison, PA. (1985). The effects of sulphur dioxide on net C02
assimilation in the lichen Evernia Mesomorpha NYL. New Phytol 100: 643-651.
http://dx.doi.Org/10.llll/i.1469-8137.1985.tb02809.x.
Hughes, BB; Ebv, R, on: Van Dyke, E; Tinker, MT, im; Marks, CI: Johnson, KS: Wasson, K.
(2013).	Recovery of a top predator mediates negative eutrophic effects on seagrass. Proc
Natl Acad Sci USA 110: 15313-15318. http://dx.doi.org/10.1073/pnas.130280511Q.
Hughes, RG: Potouroglou, M; Ziauddin, Z; Nicholls, JC. (2017). Seagrass wasting disease:
Nitrate enrichment and exposure to a herbicide (Diuron) increases susceptibility of
Zostera marina to infection. Mar Pollut Bull.
http://dx.doi.Org/10.1016/i.marpolbul.2017.08.032.
Hultberg, H; Andersson, IB. (1982). Liming of acidified lakes: Induced long-term changes.
Water Air Soil Pollut 18: 311-331. http://dx.doi.org/10.1007/BF0241942Q.
Hundev, EJ; Moser, KA: Longstaffe, FJ; Michelutti, N: Hladvniuk, R. (2014). Recent changes in
production in oligotrophic Uinta Mountain lakes, Utah, identified using paleolimnology.
Limnol Oceanogr 59: 1987-2001. http://dx.doi.Org/10.4319/lo.2014.59.6.1987.
Hundev, EJ: Russell, SD; Longstaffe, FJ: Moser, KA. (2016). Agriculture causes nitrate
fertilization of remote alpine lakes. Nat Commun 7: 10571.
http://dx.doi.org/10.1038/ncommsl0571.
Hunt, JF; Ohno, T; Fernandez, IJ. (2008). Influence of foliar phosphorus and nitrogen contents
on chemical properties of water extractable organic matter derived from fresh and
decomposed sugar maple leaves. Soil Biol Biochem 40: 1931-1939.
http://dx.doi.Org/10.1016/i.soilbio.2008.04.004.
Hurd, TM; Brach, AR; Raynal, DJ. (1998). Response of understory vegetation of Adirondack
forests to nitrogen additions. Can J For Res 28: 799-807. http://dx.doi.org/10.1139/x98-
045.
June 2018
R-102
DRAFT: Do Not Cite or Quote

-------
Hurkuck. M; Bruemmer, C; Mohr. K; Spott O: Well R; Flessa. H; Kutsch. WL. (2015). Effects
of grass species and grass growth on atmospheric nitrogen deposition to a bog ecosystem
surrounded by intensive agricultural land use. Ecol Evol 5: 2556-2571.
http://dx.doi.org/10.10Q2/ece3.1534.
Hurteau. M; North. M. (2008). Mixed-conifer understory response to climate change, nitrogen,
and fire. Global Change Biol 14: 1543-1552. http://dx.doi.Org/10.l 11 l/j.1365-
2486.2008.01584.x.
Hurteau. M; North. M. (2009). Response of Arnica dealbata to climate change, nitrogen
deposition, and fire. Plant Ecol 202: 191-194. http://dx.doi.org/10.1007/sl 1258-008-
9472-4.
Hurteau. MP: North. M; Foin. T. (2009). Modeling the influence of precipitation and nitrogen
deposition on forest understory fuel connectivity in Sierra Nevada mixed-conifer forest.
Ecol Modell 220: 2460-2468. http://dx.doi.Org/10.1016/i.ecolmodel.2009.06.032.
Hutchinson. J: Mavnard. D; Geiser. L. (1996). Air quality and lichens - a literature review
emphasizing the Pacific Northwest, USA. Washington, DC: U.S. Department of
Agriculture.
Hutchison, JS; Henry, HAL. (2010). Additive effects of warming and increased nitrogen
deposition in a temperate old field: Plant productivity and the importance of winter.
Ecosystems 13: 661-672. http://dx.doi.org/10.1007/slQ021-010-9344-3.
Hyde, EJ; Simons, TR. (2001). Sampling plethodontid salamanders: Sources of variability. J
Wildl Manag 65: 624-632.
Hyer, KE; Webb, JR; Eshleman, KN. (1995). Episodic acidification of three streams in
Shenandoah National Park, Virginia, USA. Water Air Soil Pollut 85: 523-528.
http://dx.doi.org/10.1007/BF0Q476882.
Hyvonen, R; Persson, T; Andersson, S: Olsson, B; Agren, GI; Linder, S. (2008). Impact of long-
term nitrogen addition on carbon stocks in trees and soils in northern Europe.
Biogeochemistry 89: 121-137. http://dx.doi.org/10.1007/slQ533-007-9121-3.
Ialeggio, JS: Nyman, JA. (2014). Nutria grazing preference as a function of fertilization.
Wetlands 34: 1039-1045. http://dx.doi.org/10.1007/sl3157-014-Q557-7.
Ianistcki, M; Dallarosa, J: Sauer, C: Teixeira, CE; da Silva, J. (2009). Genotoxic effect of
polycyclic aromatic hydrocarbons in the metropolitan area of Porto Alegre, Brazil,
evaluated by Helix aspersa (Miiller, 1774). Environ Pollut 157: 2037-2042.
http://dx.doi.Org/10.1016/i.envpol.2009.02.025.
Ibanez, I; Zak, PR: Burton, AJ; Pregitzer, KS. (2016). Chronic nitrogen deposition alters tree
allometric relationships: implications for biomass production and carbon storage. Ecol
Appl 26: 913-925. http://dx.doi.org/10.1890/15-0883.
Inamdar, SP; Mitchell, MJ. (2008). Sulfate exports from multiple catchments in a glaciated
forested watershed in western New York, USA. Environ Monit Assess 139: 227-245.
http://dx.doi.org/10.1007/slQ661-007-9830-z.
June 2018
R-103
DRAFT: Do Not Cite or Quote

-------
Ingersoll. CG; Gullev. DP; Mount PR: Mueller. ME; Fernandez. JD; Hockett JR: Bergman.
HL. (1990). Aluminum and acid toxicity to two strains of brook trout (Salvelinus
fontinalis). Can J Fish Aquat Sci 47: 1641-1648. http://dx.doi.Org/10.l 139/f90-l86.
Ingvorsen. K; Brock. TP. (1982). Electron flow via sulfate reduction and methanegenesis in the
anaerobic hypolimnion of Lake Mendota. Limnol Oceanogr 27: 559-564.
http://dx.doi.Org/10.4319/lo.1982.27.3.0559.
Inouve. RS. (2006). Effects of shrub removal and nitrogen addition on soil moisture in sagebrush
steppe. J Arid Environ 65: 604-618. http://dx.doi.Org/10.1016/i.iaridenv.2005.10.005.
Interlandi, SJ; Kilham, SS. (1998). Assessing the effects of nitrogen deposition on mountain
waters: A study of phytoplankton community dynamics. Water Sci Technol 38: 139-146.
http://dx.doi.org/10.1016/S0273-1223(98W)743-4.
Interlandi. SJ: Kilham. SS. (2001). Limiting resources and the regulation of diversity in
phytoplankton communities. Ecology 82: 1270-1282. http://dx.doi.org/10.1890/0Q12-
9658(200 n082ri270:LRATR012.0.CQ:2.
Interlandi. SJ: Kilham. SS: Theriot. EC. (1999). Responses of phytoplankton to varied resource
availability in large lakes of the Greater Yellowstone ecosystem. Limnology 44: 668-682.
http://dx.doi.Org/10.4319/lo.1999.44.3.0668.
IPCC (Intergovernmental Panel on Climate Change). (2007a). Climate change 2007: Impacts,
adaptation and vulnerability. Cambridge, UK: Cambridge University Press.
https://www.ipcc.ch/publications and data/publications ipcc fourth assessment report
wg2 report impacts adaptation and vulnerability.htm.
IPCC (Intergovernmental Panel on Climate Change). (2007b). Climate change 2007: The
physical science basis. New York, NY: Cambridge University Press.
http://www.ipcc.ch/publications and data/publications ipcc fourth assessment report
wgl report the physical science basis.htm.
IPCC (Intergovernmental Panel on Climate Change). (2013). Climate change 2013: The physical
science basis. Contribution of working group I to the fifth assessment report of the
Intergovernmental Panel on Climate Change. In TF Stocker; P Qin; G Plattner; MMB
Tignor; SK Allen; J Boschung; A Mauels; Y Xia; V Bex; PM Midgley (Eds.), Climate
Change 2013: The Physical Science Basis. New York, NY: Cambridge University Press.
http://www.ipcc. ch/report/ar5/wgl/#.Um6aMXPkvNk.
Irvine. IC: Greaver. T; Phelan. J: Sabo. RP; Van Houtven. G. (2017). Terrestrial acidification
and ecosystem services: effects of acid rain on bunnies, baseball, and Christmas trees.
Ecosphere 8. http://dx.doi.org/10.1002/ecs2.1857.
Irvine. IC: Vivanco. L; Bentlev. PN: Martiny. JBH. (2012). The effect of nitrogen enrichment on
C-l-cycling microorganisms and methane flux in salt marsh sediments. FMICB 3: 90.
http://dx.doi.org/10.3389/fmicb.2012.0009Q.
June 2018
R-104
DRAFT: Do Not Cite or Quote

-------
Irving. PM; Miller. JE; Xerikos. PB. (1982). The effect of N02 and S02 alone and in
combination on the productivity of field-grown soybeans. In T Schneider; L Grant (Eds.),
Studies in Environmental Science 21: Air pollution by nitrogen oxides: Proceedings of
the US-Dutch international symposium, Maastricht, The Netherlands, May 24-28, 1982
(pp. 521-531). Amsterdam, The Netherlands: Elsevier. http://dx.doi.org/10.1016/B978-Q-
444-42127-2.50052-X.
Isbell, F; Calcagno, V; Hector, A; Connolly, J; Harpole, WS; Reich, PB; Scherer-Lorenzen, M;
Schmid, B; Tilman, D; van Ruiiven, J: Weigelt, A: Wilsev, BJ; Zavaleta, ES: Loreau, M.
(2011). High plant diversity is needed to maintain ecosystem services. Nature 477: 199-
202. http://dx.doi.org/10.1038/naturel0282.
Isbell, F; Craven, D; Connolly, J; Loreau, M; Schmid, B; Beierkuhnlein, C; Bezemer, TM;
Bonin, C: Bruelheide, H; de Luca, E; Ebeling, A: Griffin, JN; Guo, Q: Hautier, Y;
Hector, A: Jentsch, A: Krevling, J: Lanta, V: Manning, P; Meyer, ST: Mori, AS: Naeem,
S: Niklaus, PA: Pollev, HW: Reich, PB: Roscher, C: Seabloom, EW: Smith, MP:
Thakur, MP: Tilman, D: Tracy, BF: van Per Putten, WH: van Ruiiven, J: Weigelt, A:
Weisser, WW: Wilsev, B: Eisenhauer, N. (2015). Biodiversity increases the resistance of
ecosystem productivity to climate extremes. Nature 526: 574-U263.
http://dx.doi.org/10.1038/naturel5374.
Isbell, F: Reich, PB: Tilman, D: Hobbie, SE: Polasky, S: Binder, S. (2013a). Nutrient
enrichment, biodiversity loss, and consequent declines in ecosystem productivity. Proc
Natl Acad Sci USA 110: 11911-11916. http://dx.doi.org/10.1073/pnas.131088011Q.
Isbell, F: Tilman, D: Polasky, S: Binder, S: Hawthorne, P. (2013b). Low biodiversity state
persists two decades after cessation of nutrient enrichment. Ecol Lett 16: 454-460.
http://dx.doi.org/10. Ill 1/ele. 12066.
Ishida, TA: Nordin, A. (2010). No evidence that nitrogen enrichment affect fungal communities
of Vaccinium roots in two contrasting boreal forest types. Soil Biol Biochem 42: 234-
243. http://dx.doi.Org/10.1016/i.soilbio.2009.10.021.
Ito, M: Mitchell, MJ: Driscoll, CT. (2002). Spatial patterns of precipitation quantity and
chemistry and air temperature in the Adirondack region of New York. Atmos Environ 36:
1051-1062. http://dx.doi.org/10.1016/S1352-2310(0n00484-8.
Iversen, CM; Bridgham, SD; Kellogg, LE. (2010). Scaling plant nitrogen use and uptake
efficiencies in response to nutrient addition in peatlands. Ecology 91: 693-707.
http://dx.doi.org/10.1890/09-0064.1.
Iwald, J: Lofgren, S: Stendahl, J: Karltun, E. (2013). Acidifying effect of removal of tree stumps
and logging residues as compared to atmospheric deposition. For Ecol Manage 290: 49-
58. http://dx.doi.Org/10.1016/i.foreco.2012.06.022.
Izuta, T: Matsumura, H: Ohashi, T: Miyake, H: Totsuka, T. (1993). Effects of peroxyacetyl
nitrate on the growth of petunia, kidney bean and radish plants. Environ Sci 2: 63-75.
Jackson, DA: Harvey, HH. (1995). Gradual reduction and extinction of fish populations in acid
lakes. Water Air Soil Pollut 85: 389-394. http://dx.doi.org/10.1007/BF0047686Q.
June 2018
R-105
DRAFT: Do Not Cite or Quote

-------
Jacobs. MI; Burke. WJ; El rod. MJ. (2014). Kinetics of the reactions of isoprene-derived
hydroxynitrates: Gas phase epoxide formation and solution phase hydrolysis. Atmos
Chem Phys 14: 8933-8946. http://dx.doi.org/10.5194/acp-14-8933-2014.
Jacob son. MZ. (2002). Atmospheric pollution: History, science, and regulation. In Atmospheric
Pollution: History, Science, and Regulation. New York: Cambridge University Press.
http://www.cambridge.org/us/academic/subiects/earth-and-environmental-
science/atmospheric-science-and-meteorology/atmospheric-pollution-hi story-science-
and-regulation.
Jaconis. SY; Cullev. TM; Meier. AM. (2017). Does particulate matter along roadsides interfere
with plant reproduction? A comparison of effects of different road types on Cichorium
intybus pollen deposition and germination. Environ Pollut 222: 261-266.
http://dx.doi.Org/10.1016/i.envpol.2016.12.047.
Jaegle. L; Steinberger. L; Martin. RV; Chance. K. (2005). Global partitioning of NOx sources
using satellite observations: Relative roles of fossil fuel combustion, biomass burning and
soil emissions. Faraday Discuss 130: 407-423. http://dx.doi.org/10.1039/b502128f.
Jager. H. -J: Klein. H. (1977). Biochemical and physiological detection of sulfur dioxide injury
to pea plants (Pisum sativum). J Air Waste Manag Assoc 27: 464-466.
http://dx.doi.org/10.1080/00022470.1977.1047Q443.
Jamieson. MA: Quintero. C: Blumenthal. DM. (2013). Interactive effects of simulated nitrogen
deposition and altered precipitation patterns on plant allelochemical concentrations. J
Chem Ecol 39: 1204-1208. http://dx.doi.org/10.1007/slQ886-013-0340-x.
Jamieson. MA: Seastedt. TR; Bowers. MP. (2012). Nitrogen enrichment differentially affects
above and belowground plant defense. Am J Bot 99: 1630-1637.
http://dx.doi.org/10.3732/aib.1100492.
Janicki Environmental, Inc., . (2011). Proposed numeric nutrient criteria for Tampa Bay. St.
Petersburg, FL: Tampa Bay Estuary Program.
https://www.tbeptech.org/attachments/article/97/TBEP Estuarine NNC Recommendatio
n Septemberl32011.pdf.
Janicki Environmental, Inc., . (2013). Estimates of total nitrogen, total phosphorus, total
suspended solids, and biochemical oxygen demand loadings to Tampa Bay, Florida:
2007-2011. St. Petersburg, FL: Tampa Bay Estuary Program.
http://www.tbeptech.org/TBEP TECH PUBS/2013/TBEP 03 13 FINAL TBEP Loads
2007-2011 %2019Mar2013 .pdf.
Janssens, IA; Dieleman, W: Luvssaert, S: Subke, J, -A: Reichstein, M; Ceulemans, R; Ciais, P;
Dolman, AJ; Grace, J: Matteucci, G: Papale, D; Piao, SL; Schulze, ED: Tang, J: Law,
BE. (2010). Reduction of forest soil respiration in response to nitrogen deposition
[Review], Nat Geosci 3: 315-322. http://dx.doi.org/10.1038/ngeo844.
Jansson, M; Bergstrom, AK; Drakare, S: Blomqvist, P. (2001). Nutrient limitation of
bacterioplankton and phytoplankton in humic lakes in northern Sweden. Freshw Biol 14:
76-85.
Jantti, H; Hietanen, S. (2012). The effects of hypoxia on sediment nitrogen cycling in the Baltic
Sea. Ambio 41: 161-169. http://dx.doi.org/10.1007/sl3280-011-Q233-6.
June 2018
R-106
DRAFT: Do Not Cite or Quote

-------
Jaramillo. P; Muller. NZ. (2016). Air pollution emissions and damages from energy production
in the US: 2002-2011. Energy Policy 90: 202-211.
http://dx.doi.Org/10.1016/i.enpol.2015.12.035.
Jarvenpaa. M; Lindstrom. K. (2011). Algal blooms decrease care but increase egg survival in a
fish with paternal care. Behav Ecol Sociobiol 65: 2023-2028.
http://dx.doi.org/10.1007/sQ0265-011-1211-5.
Jarvis. S: Woodward. S: Alexander. U; Taylor. AFS. (2013). Regional scale gradients of climate
and nitrogen deposition drive variation in ectomycorrhizal fungal communities associated
with native Scots pine. Global Change Biol 19: 1688-1696.
http://dx.doi.org/10. Ill 1/gcb. 12178.
Jassbv. AD: Reuter. JE; Axler. RP; Goldman. CR; Hacklev. SH. (1994). Atmospheric deposition
of nitrogen and phosphorus in the annual nutrient load of Lake Tahoe (California-
Nevada). Water Resour Res 30: 2207-2216. http://dx.doi.org/10.1029/94WR00754.
Jefts. S: Fernandez. IJ; Rustad. LE; Pail. DB. (2004). Decadal responses in soil N dynamics at
the Bear Brook Watershed in Maine, USA. For Ecol Manage 189: 189-205.
http://dx.doi.Org/10.1016/i.foreco.2003.08.011.
Jensen, NK; Holzmueller, EJ; Edwards, PJ; Thomas-Van Gundy, M; DeWalle, PR: Williard,
KWJ. (2014). Tree response to experimental watershed acidification. Water Air Soil
Pollut 225. http://dx.doi.org/10.1007/sll270-014-2Q34-6.
Jeppesen, E; Meerhoff, M; Davidson, TA; Trolle, D; S0ndergaard, M; Lauridsen, TL; Beklioglu,
M; Brucet, S; Volta, P; Gonzalez-Bergonzoni, I; Nielsen, A. (2014). Climate change
impacts on lakes: An integrated ecological perspective based on a multi-faceted
approach, with special focus on shallow lakes. Journal of Limnology 73: 88-111.
http://dx.doi.org/10.4081/ilimnol.2014.844.
Jeppesen, E; Meerhoff, M; Jacobsen, BA; Hansen, RS; S0ndergaard, M; Jensen, JP; Lauridsen,
TL: Mazzeo, N: Branco, CWC. (2007). Restoration of shallow lakes by nutrient control
and biomanipulation-the successful strategy varies with lake size and climate.
Hydrobiologia 581: 269-285. http://dx.doi.org/10.1007/sl0750-006-05Q7-3.
Jeppesen, E; Mehner, T; Winfield, IJ: Kangur, K; Sarvala, J: Gerdeaux, D; Rask, M; Malmquist,
HJ: Holmgren, K; Volta, P; Romo, S: Eckmann, R; Sandstrom, A: Blanco, S: Kangur, A:
Stabo, HR; Tarvainen, M; Ventela, A, -M; S0ndergaard, M; Lauridsen, TL: Meerhoff, M.
(2012). Impacts of climate warming on the long-term dynamics of key fish species in 24
European lakes. Hydrobiologia 694: 1-39. http://dx.doi.org/10.1007/slQ750-012-l 182-1.
Jeremiason, JD; Engstrom, PR: Swain, EB; Nater, EA; Johnson, BM; Almendinger, JE:
Monson, BA: Kolka, RK. (2006). Sulfate addition increases methylmercury production in
an experimental wetland. Environ Sci Technol 40: 3800-3806.
http://dx.doi.org/10.1021/es0524144.
Jeremiason, JD: Reiser, TK; Weitz, RA; Berndt, ME: Aiken, GR. (2016). Aeshnid dragonfly
larvae as bioindicators of methylmercury contamination in aquatic systems impacted by
elevated sulfate loading. Ecotoxicology 25: 456-468. http://dx.doi.org/10.1007/slQ646-
015-1603-9.
June 2018
R-107
DRAFT: Do Not Cite or Quote

-------
Jewett EB; Eldridge, PM; Burke. MK; Buxton. HT; Diaz. RJ; Lopez. CB; Kidwell. DM;
Bricker. SB; Greene. RM; Hagy. JD. Ill; Walbridge. MR. (2010). Scientific assessment
of hypoxia in U.S. coastal waters. Washington, DC: Interagency Working Group on
Harmful Algal Blooms, Hypoxia, and Human Health of the Joint Subcommittee on
Ocean Science and Technology.
https://obamawhitehouse.archives.gov/sites/default/files/microsites/ostp/hypoxia-
report.pdf.
Jeziorski. A; Keller. B; Paterson. AM; Greenawav. CM; Smol. JP. (2013). Aquatic ecosystem
responses to rapid recovery from extreme acidification and metal contamination in lakes
near Wawa, Ontario. Ecosystems 16: 209-223. http://dx.doi.org/10.1007/slQ021-012-
9603-6.
Jeziorski. A; Paterson. AM; Smol. JP. (2012a). Changes since the onset of acid deposition
among calcium-sensitive cladoceran taxa within softwater lakes of Ontario, Canada. J
Paleolimnol 48: 323-337. http://dx.doi.org/10.1007/slQ933-012-9600-v.
Jeziorski, A; Paterson, AM; Smol, JP. (2012b). Crustacean zooplankton sedimentary remains
from calcium-poor lakes: Complex responses to threshold concentrations. Aquat Sci 74:
121-131. http://dx.doi.Org/10.1007/s00027-011-0202-v.
Jeziorski, A; Paterson, AM; Watson, I; Cumming, BF; Smol, JP. (2014). The influence of
calcium decline and climate change on the cladocerans within low calcium, circumneutral
lakes of the Experimental Lakes Area. Hydrobiologia 722: 129-142.
http://dx.doi.org/10.1007/slQ750-013-1691-6.
Jeziorski, A; Yan, ND; Paterson, AM; Desellas, AM; Turner, MA; Jeffries, DS; Keller, B;
Weeber, RC; Mcnicol, D, onK; Palmer, ME; Mciver, K; Arseneau, K; Ginn, BK;
Cumming, BF; Smol, JP. (2008). The Widespread Threat of Calcium Decline in Fresh
Waters. Science 322: 1374-1377. http://dx.doi.Org/10.l 126/science.l 164949.
Jia, Y; Bao, P; Zhu, Y, -G. (2015). Arsenic bioavailability to rice plant in paddy soil: Influence
of microbial sulfate reduction. Journal of Soils and Sediments 15: 1960-1967.
http://dx.doi.org/10.1007/sll368-015-1133-3.
Jickells, T; Baker, AR; Cape, JN; Cornell, SE; Nemitz, E. (2013). The cycling of organic
nitrogen through the atmosphere [Review], Philos Trans R Soc Lond B Biol Sci 368:
20130115. http://dx.doi.org/10.1098/rstb.2013.0115.
Jimenez Cisneros, BE; Oki, T; Arnell, NW; Benito, G; Coglev, JG; Doll, P; Jiang, T; Mwakalila,
SS. (2014). Freshwater resources. In CB Field; VR Barros; DJ Dokken; KJ Mach; MD
Mastrandrea; TE Bilir; M Chatteijee; KL Ebi; YO Estrada; RC Genova; B Girma; ES
Kissel; AN Levy; S MacCracken; PR Mastrandrea; LL White (Eds.), Climate Change
2014: Impacts, Adaptation, and Vulnerability Part A: Global and Sectoral Aspects
Contribution of Working Group II to the Fifth Assessment Report of the
Intergovernmental Panel on Climate Change (pp. 229-269). Cambridge, United
Kingdom: Cambridge University Press. http://ipcc-wg2.gov/AR5/report/full-report/.
Jirousek, M; Haiek, M; Bragazza, L. (2011). Nutrient stoichiometry in Sphagnum along a
nitrogen deposition gradient in highly polluted region of Central-East Europe. Environ
Pollut 159: 585-590. http://dx.doi.Org/10.1016/i.envpol.2010.10.004.
June 2018
R-108
DRAFT: Do Not Cite or Quote

-------
Jirousek. M; Turn a. I; Zahora. J; Holub. P; Kintrova. K; Haiek. M. (2015). Decomposition
patterns, nutrient availability, species identities and vegetation changes in central-
European summit bogs. Boreal Environmental Research 20: 571-586.
Jochner. S; Markevvch. I; Beck. I; Traidl-Hoffmann. C; Heinrich. J; Menzel. A. (2015). The
effects of short- and long-term air pollutants on plant phenology and leaf characteristics.
Environ Pollut 206: 382-389. http://dx.doi.Org/10.1016/i.envpol.2015.07.040.
Johansson. O; Nordin. A: Olofsson. J: Palmqvist. K. (2010). Responses of epiphytic lichens to an
experimental whole-tree nitrogen-deposition gradient. New Phytol 188: 1075-1084.
http://dx.doi.org/10.111 l/i.l469-8137.2010.03426.x.
Johansson. O; Olofsson. J: Giesler. R; Palmqvist. K. (2011). Lichen responses to nitrogen and
phosphorus additions can be explained by the different symbiont responses. New Phytol
191: 795-805. http://dx.doi.org/10.1111/i. 1469-8137.2011,03739.x.
Johansson, O; Palmqvist, K; Olofsson, J. (2012). Nitrogen deposition drives lichen community
changes through differential species responses. Global Change Biol 18: 2626-2635.
http://dx.doi.org/10.1111/i. 1365-2486.2012.02723 ,x.
Johnes, PJ. (1996). Evaluation and management of the impact of land use change on the nitrogen
and phosphorus load delivered to surface waters: The export coefficient modelling
approach. JHydrol 183: 323-349. http://dx.doi.org/10.1016/0022-1694(95^)02951-6.
Johnson, AH: Andersen, SB: Siccama, TG. (1994a). Acid rain and soils of the Adirondacks. I.
Changes in pH and available calcium, 19301984. Can J For Res 24: 39-45.
http://dx.doi.Org/10.l 139/x94-006.
Johnson, AH: Friedland, AJ; Miller, EK; Siccama, TG. (1994b). Acid rain and soils of the
Adirondacks III Rates of soil acidification in a montane spruce-fir forest at Whiteface
Mountain, New York. Can J For Res 24: 663-669. http://dx.doi.Org/10.l 139/x94-089.
Johnson, CE; Driscoll, CT; Blum, JD; Fahev, TJ; Battles, JJ. (2014). Soil chemical dynamics
after calcium silicate addition to a northern hardwood forest. Soil Sci Soc Am J 78: 1458-
1468. http://dx.doi.org/10.2136/sssai2014.03.0114.
Johnson, D. (2001). Effects of forest management on soil C and N storage: Meta analysis. For
Ecol Manage 140: 227-238. http://dx.doi.org/10.1016/SQ378-l 127(00^)00282-6.
Johnson, DS: Warren, RS: Deegan, LA: Mozdzer, TJ. (2016a). Saltmarsh plant responses to
eutrophication. Ecol Appl 26: 2647-2659. http://dx.doi.org/10.1002/eap. 1402.
Johnson, DW: Hunsaker, CT: Glass, DW: Rau, BM; Roath, BA. (2011). Carbon and nutrient
contents in soils from the Kings River experimental watersheds, Sierra Nevada
Mountains, California. Geoderma 160: 490-502.
http://dx.doi.Org/10.1016/i.geoderma.2010.10.019.
Johnson, DW: Simonin, HA: Colquhoun, JR; Flack, FM. (1987). In situ toxicity tests of fishes in
acid waters. Biogeochemistry 3: 181-208. http://dx.doi.org/10.1007/BFQ2185192.
Johnson, J: Aherne, J: Cummins, T. (2015a). Base cation budgets under residue removal in
temperate maritime plantation forests. For Ecol Manage 343: 144-156.
http://dx.doi.Org/10.1016/i.foreco.2015.01.022.
June 2018
R-109
DRAFT: Do Not Cite or Quote

-------
Johnson. NC. (1993). Can fertilization of soil select less mutualistic mycorrhizae? Ecol Appl 3:
749-757. http://dx.doi.org/10.2307/19421Q6.
Johnson. NC; Rowland. PL; Corkidi. L; Allen. EB. (2008). Plant winners and losers during
grassland neutrophication differ in biomass allocation and mycorrhizas. Ecology 89:
2868-2878.
Johnson. NC: Rowland. PL: Corkidi. L; Egerton-Warburton, LM; Allen. EB. (2003). Nitrogen
enrichment alters mycorrhizal allocation at five mesic to semiarid grasslands. Ecology
84: 1895-1908. http://dx.doi.org/10.1890/0Q12-
9658(2003X)84ri895:NEAMAA12.0.CO;2.
Johnson. NW; Mitchell. CPJ: Engstrom. PR: Bailey. LT; Wasik. JKC: Berndt. ME. (2016b).
Methylmercury production in a chronically sulfate-impacted sub-boreal wetland. Environ
Sci Process Impacts 18: 725-734. http://dx.doi.org/10.1039/c6em00138f.
Johnson, PT; Ostfeld, RS; Keesing, F. (2015b). Frontiers in research on biodiversity and disease
[Review], Ecol Lett 18: 1119-1133. http://dx.doi.org/10.Ill 1/ele. 12479.
Johnson. PTJ; Chase. JM; Posch. KL; Hartson. RB; Gross. JA; Larson. P; Sutherland. PR:
Carpenter. SR. (2007). Aquatic eutrophication promotes pathogenic infection in
amphibians. ProcNatl Acad Sci USA 104: 15781-15786.
http://dx.doi.org/10.1073/pnas.0707763104.
Johnson. PTJ: Townsend. AR; Cleveland. CC: Glibert. PM; Howarth. RW: McKenzie. VJ:
Reimankova. E; Ward. MH. (2010). Linking environmental nutrient enrichment and
disease emergence in humans and wildlife. Ecol Appl 20: 16-29.
http://dx.doi.Org/10.1890/08-0633.l.
Johnson. RK; Angeler. PG. (2010). Tracing recovery under changing climate: Response of
phytoplankton and invertebrate assemblages to decreased acidification. J North Am
Benthol Soc 29: 1472-1490. http://dx.doi.Org/10.1899/09-171.l.
Johnson. TR. (1998). Climate change and Sierra Nevada snowpack. Santa Barbara, CA:
University of California, Santa Barbara.
Johnston. CA: Brown. TN. (2013). Water chemistry distinguishes wetland plant communities of
the Great Lakes coast. Aquat Bot 104: 111-120.
http://dx.doi.Org/10.1016/i.aquabot.2012.08.005.
Jones, L; Provins, A; Holland, M; Mills, G; Haves, F; Emmett, B; Hall, J: Sheppard, L; Smith, R;
Sutton, M; Hicks, K; Ashmore, M; Haines-Young, R; Harper-Simmonds, L. (2014). A
review and application of the evidence for nitrogen impacts on ecosystem services. 7:
76-88. http://dx.doi.Org/10.1016/i.ecoser.2013.09.001.
Jones, ME; Fenn, ME; Paine, TP. (2011). The effect of nitrogen additions on bracken fern and
its insect herbivores at sites with high and low atmospheric pollution. Arthropod-Plant
Inte 5: 163-173. http://dx.doi.org/10.1007/sl 1829-011-9125-8.
Jones, ME: Paine, TP. (2006). Petecting changes in insect herbivore communities along a
pollution gradient. Environ Pollut 143: 377-387.
http://dx.doi.Org/10.1016/i.envpol.2005.12.013.
June 2018
R-110
DRAFT: Do Not Cite or Quote

-------
Jones. ME; Paine. TP; Fenn, ME. (2008). The effect of nitrogen additions on oak foliage and
herbivore communities at sites with high and low atmospheric pollution. Environ Pollut
151: 434-442. http://dx.doi.Org/10.1016/i.envpol.2007.04.020.
Jones. ME: Paine. TP: Fenn. ME: Poth. MA. (2004). Influence of ozone and nitrogen deposition
on bark beetle activity under drought conditions. For Ecol Manage 200: 67-76.
http://dx.doi.Org/10.1016/i.foreco.2004.06.003.
Jones. RO: Chapman. SK. (2011). The roles of biotic resistance and nitrogen deposition in
regulating non-native understory plant diversity. Plant Soil 345: 257-269.
http://dx.doi.org/10.1007/sl 1104-011-0778-v.
Jordan. SJ: Stoffer. J: Nestlerode. JA. (2011). Wetlands as sinks for reactive nitrogen at
continental and global scales: A meta-analysis. Ecosystems 14: 144-155.
http://dx.doi. org/10.1007/s 10021 -010-9400-z.
Josephson, PC: Robinson, JM; Chiotti, J: Jirka, KJ; Kraft, CE. (2014). Chemical and biological
recovery from acid deposition within the Honnedaga Lake watershed, New York, USA.
Environ Monit Assess 186: 4391-4409. http://dx.doi.org/10.1007/slQ661-014-3706-9.
Joslin, JP; Kelly, JM: Van Miegroet, H. (1992). Soil chemistry and nutrition of North American
spruce-fir stands: Evidence for recent change. J Environ Qual 21: 12-30.
http://dx.doi.org/10.2134/ieal992.004724250021000100Q2x.
Jovan, S. (2008). Lichen bioindication of biodiversity, air quality, and climate: Baseline results
from monitoring in Washington, Oregon, and California. (PNW-GTR-737). Washington,
PC: US Pepartment of Agriculture.
Jovan, S: Riddell, J: Padgett, PE; Nash, TH, III. (2012). Eutrophic lichens respond to multiple
forms of N: Implications for critical levels and critical loads research. Ecol Appl 22:
1910-1922. http://dx.doi.Org/10.1890/ll-2075.l.
Jove, SB. (2012). Microbiology: A piece of the methane puzzle. Nature 491: 538-539.
http://dx.doi.org/10.1038/naturell749.
Joynt, J: Bischoff, M; Turco, R; Konopka, A; Nakatsu, CH. (2006). Microbial community
analysis of soils contaminaed with lead, chromium and petroleum hydrocarbons. Microb
Ecol 51: 209-219. http://dx.doi.org/10.1007/s00248-005-0205-Q.
Judd, KE; Likens, GE; Buso, PC: Bailey, AS. (2011). Minimal response in watershed nitrate
export to severe soil frost raises questions about nutrient dynamics in the Hubbard Brook
experimental forest. Biogeochemistry 106: 443-459. http://dx.doi.org/10.1007/slQ533-
010-9524-4.
Juice, SM; Fahev, TJ; Siccama, TG: Priscoll, CT; Penny, EG: Eagar, C: Cleavitt, NL; Minocha,
R; Richardson, AP. (2006). Response of sugar maple to calcium addition to Northern
Hardwood Forest. Ecology 87: 1267-1280. http://dx.doi.org/10.1890/0Q12-
9658(2006N)87r 1267:ROSMTC12.Q.CO;2.
Julian, P; Gu, B; Redfield, G. (2015). Comment on and reinterpretation of Gabriel et al. (2014)
'Fish mercury and surface water sulfate relationships in the Everglades protection area'.
Environ Manage 55: 1-5. http://dx.doi.org/10.1007/sQ0267-014-0377-9.
June 2018
R-lll
DRAFT: Do Not Cite or Quote

-------
Julian. P; Gu. B; Wright AL. (2016a). Mercury Stoichiometric Relationships in a Subtropical
Peatland. Water Air Soil Pollut 227. http://dx.doi.org/10.10Q7/sl 1270-016-3180-9.
Julian. P. II; Gu. B. (2015). Mercury accumulation in largemouth bass (Micropterus salmoides
Lacepede) within marsh ecosystems of the Florida Everglades, USA. Ecotoxicology 24:
202-214. http://dx.doi.org/10.1007/slQ646-014-1373-9.
Julian. P; Wright. AL: Osborne. TZ. (2016b). Iron and Sulfur Porewater and Surface Water
Biogeochemical Interactions in Subtropical Peatlands. Soil Sci Soc Am J 80: 794-802.
http://dx.doi.org/10.2136/sssai2015.ll.0418.
Jung, K; Chang, SX. (2012). Four years of simulated N and S depositions did not cause N
saturation in a mixedwood boreal forest ecosystem in the oil sands region in northern
Alberta, Canada. For Ecol Manage 280: 62-70.
http://dx.doi.Org/10.1016/i.foreco.2012.06.002.
Jung, K; Chang, SX; Ok, YS; Arshad, MA. (2013). Critical loads and H+ budgets of forest soils
affected by air pollution from oil sands mining in Alberta, Canada. Atmos Environ 69:
56-64. http://dx.doi.Org/10.1016/i.atmosenv.2012.12.010.
Jung, K; Ok, YS: Chang, SX. (2011). Sulfate adsorption properties of acid-sensitive soils in the
Athabasca oil sands region in Alberta, Canada. Chemosphere 84: 457-463.
http://dx.doi. org/10.1016/i. chemosphere. 2011.03.034.
Justic, D; Rabalais, NN; Turner, RE. (1995a). Stoichiometric nutrient balance and origin of
coastal eutrophication. Mar Pollut Bull 30: 41-46. http://dx.doi.org/10.1016/0Q25-
326X(94)00105-1.
Justic, D; Rabalais, NN: Turner, RE: Dortch, O. (1995b). Changes in nutrient structure of river-
dominated coastal waters: Stoichiometric nutrient balance and its conseqeuences. Estuar
Coast Shelf Sci 40: 339-356. http://dx.doi.org/10.1016/S0272-7714(05)80014-9.
Juutinen, S: Bubier, JL; Moore, TR. (2010). Responses of vegetation and ecosystem C02
exchange to 9 years of nutrient addition at Mer Bleue Bog. Ecosystems 13: 874-887.
http://dx.doi.org/10.1007/slQ021-010-9361-2.
Kahl, JS: Stoddard, JL: Haeuber, R; Paulsen, SG: Birnbaum, R; Devinev, FA: Webb, JR;
Dewalle, PR: Sharpe, W: Driscoll, CT; Herlihy, AT: Kellogg, JH; Murdoch, PS: Roy, K;
Webster, KE; Urquhart, NS. (2004). Have US surface waters responded to the 1990
Clean Air Act Amendments? [Review], Environ Sci Technol 38: 484A-490A.
http://dx.doi.org/10.1021/es0406861.
Kalbitz, K; Solinger, S: Park, JH; Michalzik, B; Matzner, E. (2000). Controls on the dynamics of
dissolved organic matter in soils: A review [Review], Soil Sci 165: 277-304.
http ://dx. doi .org/10.1097/00010694-200004000-00001.
Kaldorf, M; Kuhn, AJ; Schroder, WH; Hildebrandt, U; Bothe, H. (1999). Selective element
deposits in maize colonized by a heavy metal tolerance conferring arbuscular mycorrhizal
fungus. J Plant Physiol 154: 718-728. http://dx.doi.org/10.1016/S0176-1617(99)80250-8.
Kaldy, JE. (2014). Effect of temperature and nutrient manipulations on eel grass Zostera marina
L. from the Pacific Northwest, USA. Exp Mar Bio Ecol 453: 108-115.
http://dx.doi.Org/10.1016/i.iembe.2013.12.020.
June 2018
R-112
DRAFT: Do Not Cite or Quote

-------
Kaldv. JE; Brown. CA; Nelson. WG; Frazier. M. (2017). Macrophyte community response to
nitrogen loading and thermal stressors in rapidly flushed mesocosm systems. Exp Mar
Bio Ecol 497: 107-119. http://dx.doi.Org/10.1016/i.iembe.2017.09.022.
Kamp. A: de Beer. D; Nitsch. JL; Lavik. G: Stief. P. (2011). Diatoms respire nitrate to survive
dark and anoxic conditions. ProcNatl Acad Sci USA 108: 5649-5654.
http://dx.doi.org/10.1073/pnas.10157441Q8.
Kang. H; Fahev. TJ; Bae. K; Fisk. M; Sherman. RE: Yanai. RD; See. CR. (2016). Response of
forest soil respiration to nutrient addition depends on site fertility. Biogeochemistry 127:
113-124. http://dx.doi.org/10.1007/slQ533-015-0172-6.
Kang. J. inWoo: Chung. I. kKyo. (2017). The effects of eutrophication and acidification on the
ecophysiology of Ulva pertusa Kjellman. J Appl Phycol 29: 2675-2683.
http://dx.doi.org/10.1007/slQ811-017-1087-5.
Kang, P. -G; Mitchell, MJ. (2013). Bioavailability and size-fraction of dissolved organic carbon,
nitrogen, and sulfur at the Arbutus Lake watershed, Adirondack Mountains, NY.
Biogeochemistry 115: 213-234. http://dx.doi.org/10.1007/slQ533-013-9829-l.
Kang, P; Mitchell, MJ: Mayer, B; Campbell, JL. (2014). Isotopic Evidence for Determining the
Sources of Dissolved Organic Sulfur in a Forested Catchment. Environ Sci Technol 48:
11259-11267. http://dx.doi.org/10.1021/es502563n.
Kangur, K; Kangur, P; Ginter, K; Orru, K; Haldna, M; Mols, T; Kangur, A. (2013). Long-term
effects of extreme weather events and eutrophi cation on the fish community of shallow
lake Peipsi (Estonia/Russia). Journal of Limnology 72: 376-387.
http://dx.doi.org/10.4081/ilimnol.2013.e30.
Kanno, Y; Kulp, MA: Moore, SE. (2016). Recovery of native brook trout populations following
the eradication of nonnative rainbow trout in southern Appalachian Mountains streams.
North American Journal of Fisheries Management 36: 1325-1335.
http://dx.doi.org/10.1080/02755947.2016.12210Q4.
Kanno, Y; Kulp, MA: Moore, SE: Grossman, GD. (2017). Native brook trout and invasive
rainbow trout respond differently to seasonal weather variation: Spawning timing
matters. FreshwBiol 62: 868-879. http://dx.doi.org/10. Ill 1/fwb. 12906.
Kantola, T; Lyytikainen-Saarenmaa, P; Coulson, RN; Holopainen, M; Tchakerian, MP: Streett,
DA. (2016). Development of monitoring methods for Hemlock Woolly Adelgid induced
tree mortality within a Southern Appalachian landscape with inhibited access. iForest 9:
178-186. http://dx.doi.org/10.3832/iforl712-0Q8.
Kastl, E; Schloter-Hai, B; Buegger, F; Schloter, M. (2015). Impact of fertilization on the
abundance of nitrifiers and denitrifiers at the root-soil interface of plants with different
uptake strategies for nitrogen. Biol Fertil Soils 51: 57-64.
http://dx.doi.org/10.1007/s00374-014-Q948-l.
Kastowski, M; Hinderer, M; Vecsei, A. (2011). Long-term carbon burial in European lakes:
Analysis and estimate. Global Biogeochem Cycles 25.
http://dx.doi.org/10.1029/2010GB0Q3874.
June 2018
R-113
DRAFT: Do Not Cite or Quote

-------
Kaushal. SS; Groffman. PM; Band. LE; Elliott EM; Shields. CA; Kendall C. (2011). Tracking
nonpoint source nitrogen pollution in human-impacted watersheds. Environ Sci Technol
45: 8225-8232. http://dx.doi.org/10.1021/es20Q779e.
Kaushal. SS: Groffman. PM: Band. LE: Shields. CA: Morgan. RP; Palmer. MA: Belt. KT; Swan.
CM; Findlav. SEG: Fisher. GT. (2008). Interaction between urbanization and climate
variability amplifies watershed nitrate export in Maryland. Environ Sci Technol 42:
5872-5878. http://dx.doi.org/10.1021/es800264f.
Kaushal. SS; Likens. GE; Utz. RM; Pace. ML; Grese. M; Yepsen. M. (2013). Increased river
alkalinization in the eastern U.S. Environ Sci Technol 47: 10302-10311.
http://dx.doi.org/10.1021/es401046s.
Kaushal. SS; Mayer. PM; Vidon. PG; Smith. RM; Pennino. MJ; Newcomer. TA; Duan. S;
Weltv. C; Belt. KT. (2014). Land use and climate variability amplify carbon, nutrient,
and contaminant pulses: A review with management implications. J Am Water Resour
Assoc 50: 585-614. http://dx.doi.org/10.Ill 1/iawr. 12204.
Kaylor. SD; Hughes. MJ; Franklin. JA. (2017). Recovery trends and predictions of Fraser fir
(Abies fraseri) dynamics in the Southern Appalachian Mountains. Can J For Res 47: 125-
133. http://dx.doi.Org/10.l 139/cifr-2016-0067.
Keeler. BL; Hobbie. SE; Kellogg. LE. (2009). Effects of long-term nitrogen addition on
microbial enzyme activity in eight forested and grassland sites: Implications for litter and
soil organic matter decomposition. Ecosystems 12: 1-15.
http ://dx. doi .org/10.1007/s 10021 -008-9199-z.
Keglev. S; Orme. S; Neumeister. L. (2000). Hooked on poison: Pesticide use in California, 1991-
1998. San Francisco, CA: Californians for Pesticide Reform, Pesticide Action Network.
Keller. W. (2007). Implications of climate warming for Boreal Shield lakes: a review and
synthesis. Environ Rev 15: 99-112. http://dx.doi.Org/10.l 139/A07-002.
Keller. W; Gunn. JM. (1995). Lake water quality improvements and recovering aquatic
communities. In JM Gunn (Ed.), Restoration and Recovery of an Industrial Region (pp.
67-80). New York, NY: Springer-Verlag. http://dx.doi.org/10.1007/978-l-4612-252Q-
1 5.
Keller. W; Pitblado. TR. (1986). Water quality changes in Sudbury area lakes: A comparison of
synoptic surveys in 1974-1976 and 1981-1983. Water Air Soil Pollut 29: 285-296.
http://dx.doi.org/10.1007/BF0015876Q.
Kellev. AM; Epstein. HE. (2009). Effects of nitrogen fertilization on plant communities of
nonsorted circles in moist nonacidic tundra, Northern Alaska. Arctic Antarct Alpine Res
41: 119-127. http://dx.doi.Org/10.1657/1523-0430-41.l.119.
Kellogg. ML; Smyth. AR; Luckenbach. MW; Carmichael. RH; Brown. BL; Cornwell. JC;
Piehler. MF; Owens. MS; Dalrymple. DJ; Higgins. CB. (2014). Use of oysters to mitigate
eutrophication in coastal waters. Estuar Coast Shelf Sci 151: 156-168.
http://dx.doi.Org/10.1016/i.ecss.2014.09.025.
Kelly. CA; Rudd. JWM. (1984). Epilimnetic sulfate reduction and its relationship to lake
acidification. Biogeochemistry 1: 63-77. http://dx.doi.org/10.1007/BFQ2181121.
June 2018
R-114
DRAFT: Do Not Cite or Quote

-------
Kelly. CA; Rudd, JWM; Cook. RB; Schindler, DW. (1982). The potential importance of
bacterial processes in regulating rate of lake acidification. Limnology 27: 868-882.
http://dx.doi.Org/10.4319/lo.1982.27.5.0868.
Kelly. CA: Rudd. JWM: Hesslein. RH; Driscoll. CT; Gherini. SA; Hecky. RE. (1987). Prediction
of biological acid neutralization in acid-sensitive lakes. Biogeochemistry 3: 129-140.
Kelly. JT: Lerner. DT: O'Dea. MF: Regish. AM: Monette. MY: Hawkes. JP: Nislow. KH:
McCormick. SD. (2015). Evidence for episodic acidification effects on migrating
Atlantic salmon Salmo salar smolts. J Fish Biol 87: 1129-1146.
http://dx.doi.org/10. Ill 1/ifb. 12763.
Kelso. JRM: Jeffries. DS. (1988). Response of headwater lakes to varying atmospheric
deposition in north-central Ontario, 197985. Can J Fish Aquat Sci 45: 1905-1911.
http://dx.doi.Org/10.l 139/f88-222.
Kemp. WM: Boynton. WR. (1984). SPATIAL AND TEMPORAL COUPLING OF NUTRIENT
INPUTS TO ESTUARINE PRIMARY PRODUCTION - THE ROLE OF
PARTICULATE TRANSPORT AND DECOMPOSITION. Bulletin of Marine Science
35: 522-535.
Kemp, WM; Boynton, WR: Adolf, JE; Boesch, DF; Boicourt, WC; Brush, G; Cornwell, JC;
Fisher, TR: Glibert, PM: Hagy, JD: Harding, LW: Houde, ED: Kimmel, DG: Miller, WD:
Newell, RIE: Roman, MR: Smith, EM: Stevenson, JC. (2005). Eutrophication of
Chesapeake Bay: Historical trends and ecological interactions. Mar Ecol Prog Ser 303: 1-
29. http://dx.doi.org/10.3354/meps303001.
Kenkel, JA: Sisk, TP: Hultine, KR: Sesnie, SE: Bowker, MA: Johnson, NC. (2016). Indicators
of vehicular emission inputs into semi-arid roadside ecosystems. J Arid Environ 134:
150-159. http://dx.doi.Org/10.1016/i.iaridenv.2016.06.007.
Kennedy, CJ: Picard, C. (2012). Chronic low pH exposure affects the seawater readiness of
juvenile Pacific sockeye salmon. Fish Physiol Biochem 38: 1131-1143.
http://dx.doi.org/10.1007/slQ695-011-9599-4.
Kennison, RL: Fong, P. (2014). Extreme eutrophi cation in shallow estuaries and lagoons of
California is driven by a unique combination of local watershed modifications that trump
variability associated with wet and dry seasons. Estuaries Coasts 37: 164-179.
http://dx.doi.org/10.1007/sl2237-013-9687-z.
Kennison, RL: Kamer, K: Fong, P. (2011). Rapid nitrate uptake rates and large short-term
storage capacities may explain why opportunistic green macroalgae dominate shallow
eutrophic estuaries. J Phycol 47: 483-494. http://dx.doi.org/10. Ill 1/i .1529-
8817.2011.00994.x.
Keppel, AG: Breitburg, PL: Burrell, RB. (2016). Effects of Co-Varying Diel-Cycling Hypoxia
and pH on Growth in the Juvenile Eastern Oyster, Crassostrea virginica. PLoS ONE 11:
e0161088. http://dx.doi.org/10.1371/iournal.pone.0161Q88.
Kerin, EJ: Gilmour, CC: Roden, E: Suzuki, MT: Coates, JP: Mason, RP. (2006). Mercury
methylation by dissimilatory iron-reducing bacteria. Appl Environ Microbiol 72: 7919-
7921. http://dx.doi.org/10.1128/AEM.01602-06.
June 2018
R-115
DRAFT: Do Not Cite or Quote

-------
Kerry. A; Welbourn. PM; Prucha. B; Mierle. G. (1991). Mercury methylation by sulfate-
reducing bacterua from sediments of an acid stressed lake. Water Air Soil Pollut 56: 565-
575. http://dx.doi.org/10.1007/BF0034230Q.
Kerstiens. G: Federholzner. R: Lendzian. KJ. (1992). Dry deposition and cuticular uptake of
pollutant gases. Agric Ecosyst Environ 42: 239-253. http://dx.doi.org/10.1016/Q167-
8809(92N)90002-S.
Ket WA: Schubauer-Beri gan. JP: Craft CB. (2011). Effects of five years of nitrogen and
phosphorus additions on a Zizaniopsis miliacea tidal freshwater marsh. Aquat Bot 95: 17-
23. http://dx.doi.Org/10.1016/i.aquabot.2011.03.003.
Khalili. MI; Temnerud. J: F rob erg. M; Karl tun. E; Wevhenmever. GA. (2010). Nitrogen and
carbon interactions between boreal soils and lakes. Global Biogeochem Cycles 24.
http://dx.doi.org/10.1029/2009GB0Q3668.
Kiikkila, O. (2003). Heavy-metal pollution and remediation of forest soil around the Haijavalta
Cu-Ni smelter, in SW Finland. SilvaFennica 37: 399-415.
http://dx.doi.org/10.14214/sf.497.
Kim. CG; Bell. JNB; Power. SA. (2003). Effects of soil cadmium on Pinus sylvestris L.
seedlings. Plant Soil 257: 443-449. http://dx.doi.Org/10.1023/A:1027380507087.
Kim. I. -N; Lee. K; Gruber. N; Karl. DM; Bullister. JL; Yang. S; Kim. T. -W. (2014a). Chemical
oceanography. Increasing anthropogenic nitrogen in the North Pacific Ocean. Science
346: 1102-1106. http://dx.doi.org/10.1126/science. 1258396.
Kim. J. -S; Jain. S; Norton. SA. (2010). Streamflow variability and hydroclimatic change at the
Bear Brook Watershed in Maine (BBWM), USA. Environ Monit Assess 171: 47-58.
http://dx.doi.org/10.1007/slQ661-010-1525-l.
Kim. PS; Jacob. DJ; Fisher. JA; Travis. K; Yu. K; Zhu. L; Yantosca. RM; Sulprizio. MP;
Jimenez. JL; Campuzano-Jost. P; Frovd. KD; Liao. J; Hair. JW; Fenn. MA; Butler. CF;
Wagner. NL; Gordon. TP; Welti. A; Wennberg. PO; Crounse. JD; St. Clair. JM; Teng.
AP; Millet. DB; Schwarz. JP; Markovic. MZ; Perring. AE. (2015a). Sources, seasonality,
and trends of southeast US aerosol: An integrated analysis of surface, aircraft, and
satellite observations with the GEOS-Chem chemical transport model. Atmos Chem Phys
15: 10411-10433. http://dx.doi.org/10.5194/acp-15-10411-2015.
Kim. SY; Veraart. AJ; Meima-Franke. M; Bodelier. PLE. (2015b). Combined effects of carbon,
nitrogen and phosphorus on CH4 production and denitrification in wetland sediments.
Geoderma259-260: 354-361. http://dx.doi.Org/10.1016/i.geoderma.2015.03.015.
Kim. T. -W; Najjar. RG; Lee. K. (2014b). Influence of precipitation events on phytoplankton
biomass in coastal waters of the eastern United States. Global Biogeochem Cycles 28: 1-
13. http://dx.doi.org/10.1002/2013GB0Q4712.
Kim. T; Lee. K; Najjar. RG; Jeong. H; Jeong. H. (2011). Increasing N Abundance in the
Northwestern Pacific Ocean Due to Atmospheric Nitrogen Deposition. Science 334: 505-
509. http://dx.doi.org/10.1126/science. 1206583.
June 2018
R-116
DRAFT: Do Not Cite or Quote

-------
Kimball S; Goulden. ML; Suding, KN; Parker. S. (2014). Altered water and nitrogen input shifts
succession in a southern California coastal sage community. Ecol Appl 24: 1390-1404.
http://dx.doi.ore/10.1890/13-1313.1.
King. GM; Wiebe, WJ. (1978). Methane release from soils of a Georgia salt marsh. Geochim
Cosmo Act 42: 343-348. http://dx.doi.org/10.1016/0016-7037(78^)90264-8.
King. JK; Saunders. FM; Lee. RF; Jahnke. RA. (1999). Coupling mercury methylation rates to
sulfate reduction rates in marine sediments. Environ Toxicol Chem 18: 1362-1369.
http://dx.doi.org/10.1002/etc.56201807Q4.
Kirby, CS; Mclnerney, B; Turner, MP. (2008). Groundtruthing and potential for predicting acid
deposition impacts in headwater streams using bedrock geology, GIS, angling, and
stream chemistry. Sci Total Environ 393: 249-261.
http://dx.doi.Org/10.1016/i.scitotenv.2007.12.026.
Kj0ller, R; Nilsson, LO; Hansen, K; Schmidt, IK; Vesterdal, L; Gundersen, P. (2012). Dramatic
changes in ectomycorrhizal community composition, root tip abundance and mycelial
production along a stand-scale nitrogen deposition gradient. New Phytol 194: 278-286.
http://dx.doi.Org/10.llll/i.1469-8137.2011.04041.x.
Klauda, RJ; Palmer, RE. (1987). Responses of blueback herring eggs and larvae to pulses of acid
and aluminum. Trans Am Fish Soc 116: 561-569. http://dx.doi.org/10.1577/1548-
8659(1987^)116<561 :ROBHEA>2.Q.CO:2.
Kleeberg, A; Neven, M; Kalettka, T. (2016). Element-specific downward fluxes impact the
metabolism and vegetation of kettle holes. Hydrobiologia 766: 261-274.
http://dx.doi.org/10.1007/slQ750-015-2460-5.
Kleijn, D; Bekker, RM; Bobbink, R; De Graaf, MCC; Roelofs, JGM. (2008). In search for key
biogeochemical factors affecting plant species persistence in heathland and acidic
grasslands: a comparison of common and rare species. J Appl Ecol 45: 680-687.
http://dx.doi.org/10.111 l/i.l365-2664.2007.01444.x.
Kleindienst, TE. (1994). Recent developments in the chemistry and biology of peroxyacetyl
nitrate. Res Chem Intermed 20: 335-384. http://dx.doi.Org/10.l 163/156856794X00379.
Klemas, V. (2012). Remote Sensing of Algal Blooms: An Overview with Case Studies. J Coast
Res 28: 34-43. http://dx.doi.org/10.2112/JCOASTRES-D-l 1-00051.1.
Kline, KM; Eshleman, KN; Garlitz, JE; U'Ren, SH. (2016). Long-term response of surface water
acid neutralizing capacity in a central Appalachian (USA) river basin to declining acid
deposition. Atmos Environ 146: 195-205.
http://dx.doi.Org/10.1016/i.atmosenv.2016.07.034.
Klopatek, JM; Barry, MJ; Johnson, DW. (2006). Potential canopy interception of nitrogen in the
Pacific Northwest, USA. For Ecol Manage 234: 344-354.
http://dx.doi.Org/10.1016/i.foreco.2006.07.019.
Knapp, A, k; Briggs, JM; Hartnett, DC; Collins, SL. (1998). Grassland dynamics: Long-term
ecological research in tallgrass prairie. In A Knapp; JM Briggs; DC Hartnett; SL Collins
(Eds.), Grassland Dynamics: Long-Term Ecological Research in Tallgrass Prairie (1st
ed.). London, UK: Oxford University Press.
June 2018
R-117
DRAFT: Do Not Cite or Quote

-------
Kneitel. JM; Lessin. CL. (2010). Ecosystem-phase interactions: aquatic eutrophication decreases
terrestrial plant diversity in California vernal pools. Oecologia 163: 461-469.
http://dx.doi.org/10.1007/s00442-009-1529-Q.
Knoepp. JD; Vose. JM: Jackson. WA; Elliott KJ; Zarnoch. S. (2016). High elevation watersheds
in the southern Appalachians: Indicators of sensitivity to acidic deposition and the
potential for restoration through liming. For Ecol Manage 377: 101-117.
http://dx.doi.ore/10.1016/i.foreco.2016.06.040.
Knops. JMH; Nash. TH. Ill: Schlesinger. WH. (1996). The influence of epiphytic lichens on the
nutrient cycling of an oak woodland. Ecol Monogr 66: 159-179.
http://dx.doi.org/10.2307/2963473.
Knops. JMH: Reinhart. K. (2000). Specific leaf area along a nitrogen fertilization gradient. Am
Midi Nat 144: 265-272. http://dx.doi.org/10.1674/00Q3-
0031(2000N)144r0265:SLAAAN12.0.CQ:2.
Knorr. M; Frev. SD; Curtis. PS. (2005). Nitrogen additions and litter decomposition: A meta-
analysis. Ecology 86: 3252-3257. http://dx.doi.org/10.1890/05-015Q.
Kobe. RK; Likens. GE; Eagar. C. (2002). Tree seedling growth and mortality responses to
manipulations of calcium and aluminum in a northern hardwood forest. Can J For Res 32:
954-966. http://dx.doi.org/10.1139/X02-018.
Kobvlinski. A: Fredeen. AL. (2015). Importance of arboreal cyanolichen abundance to nitrogen
cycling in sub-boreal spruce and fir forests of central British Columbia, Canada. 6: 2588-
2607. http://dx.doi.org/10.3390/f6082588.
Koch. M; Bowes. G: Ross. C: Zhang. XH. ai. (2013). Climate change and ocean acidification
effects on seagrasses and marine macroalgae. Global Change Biol 19: 103-132.
http://dx.doi.Org/10.llll/i.1365-2486.2012.02791.x.
Koch. MS: Mendelssohn. IA; McKee. KL. (1990). Mechanism for the hydrogen sulfide-induced
growth limitation in wetland macrophytes. Limnology 35: 399-408.
http://dx.doi.Org/10.4319/lo.1990.35.2.0399.
Kochy. M; Wilson. SD. (2001). Nitrogen deposition and forest expansion in the Northern Great
Plains. J Ecol 89: 807-817. http://dx.doi.Org/10.1046/i.0022-0477.2001.00600.x.
Koerselman. W. (1989). Groundwater and surface water hydrology of a small groundwater-fed
fen. Wetlands Ecology and Management 1: 31-43. http://dx.doi.org/10.1007/bf00177888.
Koerselman. W: Van Kerkhoven. MB: Verhoeven. JTA. (1993). Release of inorganic N, P, and
K in peat soils; effect of temperature, water chemistry and water level. Biogeochemistry
20: 63-81. http ://dx. doi .org/10.1007/BF00004135.
Kok. JWK; Yeo. DC J: Leong. SCY. (2015). Growth and physiological responses of a tropical
toxic marine microgalga Heterosigma akashiwo (Heterokontophyta: Raphidophyceae)
from Singapore waters to varying nitrogen sources and light conditions. Ocean Sci J 50:
491-508. http://dx.doi.org/10.1007/sl2601-015-0Q45-x.
Kominoski. JS: Rosemond. A: Benstead. JP; Gulis. V: Maerz. JC: Manning. DWP. (2015). Low-
to-moderate nitrogen and phosphorus concentrations accelerate microbially driven litter
breakdown rates. Ecol Appl 25: 856-865. http://dx.doi.org/10.1890/14-l 113.1.sm.
June 2018
R-118
DRAFT: Do Not Cite or Quote

-------
Komoroske. LM; Lewi son. RL; S em in off. JA; Dehevn. DP; Dutton. PH. (2011). Pollutants and
the health of green sea turtles resident to an urbanized estuary in San Diego, CA.
Chemosphere 84: 544-552. http ://dx.doi.org/10.1016/j.chemosphere.2011.04.023.
Konneke. M; Bern hard. AE; de la Torre. JR; Walker. CB; Waterbury. JB; Stahl. DA. (2005).
Isolation of an autotrophic ammonia-oxidizing marine archaeon. Nature 437: 543-546.
http://dx.doi.org/10.1038/nature03911.
Koo. B; Kumar. N: Knipping. E; Nopmongcol. U; Sakulvanontvittava. T; Odman. MT: Russell
AG: Yarwood. G. (2015). Chemical transport model consistency in simulating regulatory
outcomes and the relationship to model performance. Atmos Environ 116: 159-171.
http://dx.doi.Org/10.1016/i.atmosenv.2015.06.036.
Koo. B; Piyachaturawat P; Morris. R; Knipping. E. (2012). Evaluation of the variability in
chemical transport model performance for deposition and ambient concentrations of
nitrogen and sulfur compounds. Atmosphere (Basel) 3: 400-418.
http://dx.doi.org/10.3390/atmos303040Q.
Koo. KA. h; Patten. BC: Teskev. RO: Creed. IF. (2014). Climate change effects on red spruce
decline mitigated by reduction in air pollution within its shrinking habitat range. Ecol
Modell 293: 81-90. http://dx.doi.Org/10.1016/i.ecolmodel.2014.07.017.
Kopacek. J: Cosby. BJ; Evans. CD: Hruska. J: Moldan. F; Oulehle. F; Santruckova. H;
Tahovska. K; Wright. RF. (2013). Nitrogen, organic carbon and sulphur cycling in
terrestrial ecosystems: Linking nitrogen saturation to carbon limitation of soil microbial
processes. Biogeochemistry 115: 33-51. http://dx.doi.org/10.1007/slQ533-013-9892-7.
Koricheva. J: Larsson. S: Haukioia. E; Keinanen. M. (1998). Regulation of woody plant
secondary metabolism by resource availability: Hypothesis testing by means of meta-
analysis. Oikos 83: 212-226. http://dx.doi.org/10.2307/3546833.
Korosi. JB; Smol. JP. (2012). Examining the effects of climate change, acidic deposition, and
copper sulphate poisoning on long-term changes in cladoceran assemblages. Aquat Sci
74: 781-792. http://dx.doi.org/10.1007/s00027-012-0261 -8.
Korthals. ET; Winfrey. MR. (1987). Seasonal and spatial variations in mercury methylation and
demethylation in an oligotrophic lake. Appl Environ Microbiol 53: 2397-2404.
Koseva. IS: Watmough. SA; Aherne. J. (2010). Estimating base cation weathering rates in
Canadian forest soils using a simple texture-based model. Biogeochemistry 101: 183-196.
http://dx.doi.org/10.1007/sl0533-010-95Q6-6.
Kosten. S: Huszar. VLM; Becares. E; Costa. LS: van Donk. E; Hansson. L. -A: Jeppesenk. E;
Kruk. C: Lacerot. G: Mazzeo. N: De Meester. L; Moss. B; Ltirling. M; Noges. T; Romo.
S: Scheffer. M. (2012). Warmer climates boost cyanobacterial dominance in shallow
lakes. Global Change Biol 18: 118-126. http://dx.doi.Org/10.l 11 l/j.1365-
2486.2011.02488.x.
Kou. L; Guo. D; Yang. H; Gao. W: Li. S. (2015). Growth, morphological traits and mycorrhizal
colonization of fine roots respond differently to nitrogen addition in a slash pine
plantation in subtropical China. Plant Soil 391: 207-218.
http://dx.doi.org/10.1007/slllQ4-015-2420-x.
June 2018
R-119
DRAFT: Do Not Cite or Quote

-------
Kou, L; Mccormack, ML; Chen. W; Guo, D: Wang. H; Gao, W; Yang. H, ao; Li. S. (2017).
Nitrogen ion form and spatio-temporal variation in root distribution mediate nitrogen
effects on lifespan of ectomycorrhizal roots. Plant Soil 411: 261-273.
http://dx.doi.org/10.1007/sl 1104-016-3018-7.
Kovats. N: Acs. A: Kovacs. A: Ferincz. A: Turoczi. B; Gelencser. A. (2012). Direct contact test
for estimating the ecotoxicity of aerosol samples. Environ Toxicol Pharmacol 33: 284-
287. http://dx.doi.Org/10.1016/i.etap.2011.12.021.
Kowalchuk. GA; Stephen. JR. (2001). Ammonia-oxidizing bacteria: A model for molecular
microbial ecology [Review], Annu Rev Microbiol 55: 485-529.
http://dx.doi.Org/10.l 146/annurev.micro.55.1.485.
Kowalik. RA; Cooper. DM: Evans. CD: Ormerod. SJ. (2007). Acidic episodes retard the
biological recovery of upland British streams from chronic acidification. Global Change
Biol 13: 2439-2452. http://dx.doi.Org/10.l 111/i. 1365-2486.2007.01437.X.
Kozlov. MY: Eranen. JK; Zverev. YE. (2007). Budburst phenology of white birch in industrially
polluted areas. Environ Pollut 148: 125-131.
http://dx.doi.Org/10.1016/i.envpol.2006.10.038.
Kraft, B; Tegetmeyer, HE; Sharma, R; Klotz, MG: Ferdelman, TG; Hettich, RL; Geelhoed, JS;
Strous. M. (2014). Nitrogen cycling. The environmental controls that govern the end
product of bacterial nitrate respiration. Science 345: 676-679.
http://dx.doi.org/10.1126/science. 1254070.
Kratz, KW: Cooper, SD; Melack, JM. (1994). Effects of single and repeated experimental acid
pulses on invertebrates in a high altitude Sierra Nevada stream. Freshw Biol 32: 161-183.
http://dx.doi.Org/10.llll/i.1365-2427.1994.tb00875.x.
Kress, LW: Skelly, JM. (1982). Response of several eastern forest tree species to chronic doses
of ozone and nitrogen dioxide. Plant Dis 66: 1149-1152. http://dx.doi.org/10.1094/PD-
66-1149.
Kretser, JN; Gallagher, WJ. (1989). Adirondack lakes study 19841987: An evaluation of fish
communities and water chemistry. Ray Brook, NY: Adirondack Lakes Survey
Corporation, http://www.adirondacklakessurvey.org/.
Kreutzer, K; Butterbach-Bahl, K; Rennenberg, H: Papen, H. (2009). The complete nitrogen cycle
of an N-saturated spruce forest ecosystem [Review], Plant Biol (Stuttg) 11: 643-649.
http://dx.doi.org/10.1111/i. 1438-8677.2009.00236.X.
Krieger, DJ. (2001). The economic value of forest ecosystem services: A review. Washington,
D.C.: The Wilderness Society.
Kroeker, KJ; Kordas, RL: Crim, R; Hendriks, IE: Ramaio, L; Singh, GS: Duarte, CM: Gattuso,
JP. (2013). Impacts of ocean acidification on marine organisms: Quantifying sensitivities
and interaction with warming. Global Change Biol 19: 1884-1896.
http://dx.doi.org/10. Ill 1/gcb. 12179.
June 2018
R-120
DRAFT: Do Not Cite or Quote

-------
Kroglund, F; Finstad, B; Stefansson, SO; Nilsen, TO; Kristensen, T; Rosseland, BO; Teien, HC;
Salbu, B. (2007). Exposure to moderate acid water and aluminum reduces Atlantic
salmon post-smolt survival. Aquaculture 273: 360-373.
http: //dx. doi. or g/10.1016/i. aquaculture .2007.10.018.
Kroglund, F; Rosseland, BO; Teien, H, -C; Salbu, B; Kristensen, T; Finstad, B. (2008). Water
quality limits for Atlantic salmon (Salmo salar L.) exposed to short term reductions in pH
and increased aluminum simulating episodes. Hydrol Earth Syst Sci 12: 491-507.
http ://dx. doi. org/10.5194/hess-12-491-2008.
Kronberg. RM; Jiskra. M; Wiederhold. JG; Biorn. E; Skyllberg. U. (2016). Methyl Mercury
Formation in Hillslope Soils of Boreal Forests: The Role of Forest Harvest and
Anaerobic Microbes. Environ Sci Technol 50: 9177-9186.
http://dx.doi.org/10.1021/acs.est.6b00762.
Kropff MJ. (1991). Long-term effects of S02 on plants, S02 metabolism and regulation of
intracellular pH. Plant Soil 131: 235-245. http://dx.doi.org/10.1007/BF000Q9454.
Krotkov. NA; McClure. B; Dickerson. RR; Cam. SA; Li. C; Bhartia. PK; Yang. K; Krueger. AJ;
Li. ZQ; Levelt. PF; Chen. HB; Wang. PC; Lu. DR. (2008). Validation of S02 retrievals
from the Ozone Monitoring Instrument over NE China. J Geophys Res Atmos 113.
http://dx.doi.org/10.1029/2007id0Q8818.
Krotkov. NA; McLinden. CA; Li. C; Lamsal. LN; Celarier. EA; Marchenko. SV; Swartz. WH;
Bucsela. EJ; Joiner. J; Duncan. BN; Boersma. KF; Veefkind. JP; Levelt. PF; Fioletov.
YE; Dickerson. RR; He. H; Lu. Z; Streets. DG. (2016). Aura OMI observations of
regional S02 and N02 pollution changes from 2005 to 2015. Atmos Chem Phys 16:
4605-4629. http://dx.doi.org/10.5194/acp-16-4605-2016.
Krug. EC; Frink. CR. (1983). Acid rain on acid soil: A new perspective. Science 221: 520-525.
http://dx.doi.Org/10.l 126/science.221.4610.520.
Kruit. RJW; van Pul. WAJ; Sauter. FJ; van Den Broek. M; Nemitz. E; Sutton. MA; Krol. M;
Holtslag. AAM. (2010). Modeling the surface-atmosphere exchange of ammonia. Atmos
Environ 44: 945-957. http://dx.doi.Org/10.1016/i.atmosenv.2009.l 1.049.
Krumins, JA; Dighton, J; Gray, D; Franklin, RB; Morin, PJ; Roberts, MS. (2009). Soil microbial
community response to nitrogen enrichment in two scrub oak forests. For Ecol Manage
258: 1383-1390. http://dx.doi.Org/10.1016/i.foreco.2009.06.046.
Krupa, S; Booker, F; Bowersox, V; Lehmann, CT; Grantz, D. (2008). Uncertainties in the
current knowledge of some atmospheric trace gases associated with U.S. agriculture: A
review [Review], J Air Waste Manag Assoc 58: 986-993. http://dx.doi.org/10.3155/lQ47-
3289.58.8.986.
Krupa, SV. (2003). Effects of atmospheric ammonia (NH3) on terrestrial vegetation: A review
[Review], Environ Pollut 124: 179-221. http://dx.doi.org/10.1016/S0269-7491(02)00434-
7.
Krynak, KL; Burke, DJ; Benard, MF. (2015). Larval Environment Alters Amphibian Immune
Defenses Differentially across Life Stages and Populations. PLoS ONE 10: e0130383.
http://dx.doi.org/10.1371/iournal.pone.0130383.
June 2018
R-121
DRAFT: Do Not Cite or Quote

-------
Krywult M; Bytnerowi cz. A. (1997). Induction of nitrate reductase activity by nitric acid vapor
in California black oak (Quercus kelloggii), canyon live oak (Quercus chrysolepis), and
ponderosa pine (Pinus ponderosa) seedlings. Can J For Res 27: 2101-2104.
http://dx.doi.Org/10.l 139/cifr-27-12-2101.
Krzyzanowski. J. (2011). Approaching cumulative effects through air pollution modelling. Water
Air Soil Pollut 214: 253-273. http://dx.doi.org/10.1007/s 11270-010-0421 -1.
Krzyzanowski. J: Innes. JL. (2010). Back to the basics - estimating the sensitivity of freshwater
to acidification using traditional approaches. J Environ Manage 91: 1227-1236.
http://dx.doi.Org/10.1016/i.ienvman.2010.01.013.
Kucharzyk. KH; Deshusses. MA: Porter. KA; Hsu-Kim. H. (2015). Relative contributions of
mercury bioavailability and microbial growth rate on net methylmercury production by
anaerobic mixed cultures. Environ Sci Process Impacts 17: 1568-1577.
http://dx.doi.org/10.1039/c5em00174a.
Kudela. RM. (2011). Characterization and deployment of solid phase adsorption toxin tracking
(SPATT) resin for monitoring of microcystins in fresh and saltwater. Harmful Algae 11:
117-125. http://dx.doi.Org/10.1016/i.hal.2011.08.006.
Kuhnel, R; Roberts, TJ; Bjorkman, MP; Isaksson, E; Aas, W; Holmen, K, im; Strom, J. (2011).
20-Year Climatology of N03- and NH4+ Wet Deposition at Ny-Alesund, Svalbard.
Advances in Meteorology. http://dx.doi.Org/10.l 155/2011/406508.
Kulkarni, MY: Burgin, A: Groffman, PM; Yavitt, JB. (2014). Direct flux and N-15 tracer
methods for measuring denitrification in forest soils. Biogeochemistry 117: 359-373.
http://dx.doi.org/10.1007/slQ533-013-9876-7.
Kunito, T; Isomura, I: Sumi, H; Park, H, oD; Toda, H; Otsuka, S: Nagaoka, K; Saeki, K; Senoo,
K. (2016). Aluminum and acidity suppress microbial activity and biomass in acidic forest
soils. Soil Biol Biochem 97: 23-30. http://dx.doi.Org/10.1016/i.soilbio.2016.02.019.
Kunza, LA: Hall, RO, Jr. (2013). Demographic and mutualistic responses of stream nitrogen
fixers to nutrients. Freshw Sci 32: 991-1004. http://dx.doi.Org/10.1899/12-102.l.
Kunza, LA: Hall, RO, Jr. (2014). Nitrogen fixation can exceed inorganic nitrogen uptake fluxes
in oligotrophic streams. Biogeochemistry 121: 537-549.
http://dx.doi.org/10.1007/slQ533-014-0021-z.
Kurek, J: Cwynar, L, esC; Weeber, RC; Jeffries, DS; Smol, JP. (2010a). Ecological distributions
of Chaoborus species in small, shallow lakes from the Canadian Boreal Shield ecozone.
Hydrobiologia 652: 207-221. http://dx.doi.org/10.1007/sl0750-010-Q333-5.
Kurek, J: Korosi, JB: Jeziorski, A: Smol, JP. (2010b). Establishing reliable minimum count sizes
for cladoceran subfossils sampled from lake sediments [journal article], 44: 603-612.
https://doi.org/10.1007/slQ933-010-9440-6.
Kusiima, JM; Powers, SE. (2010). Monetary value of the environmental and health externalities
associated with production of ethanol from biomass feedstocks. Energy Policy 38: 2785-
2796. http://dx.doi.Org/10.1016/i.enpol.2010.01.010.
June 2018
R-122
DRAFT: Do Not Cite or Quote

-------
Labai. AL; Jeziorski. A; Kurek. J; Smol. JP. (2014). Long-term trends in Cladoceran
assemblages related to acidification and subsequent liming of middle lake (Sudbury,
Canada). Water Air Soil Pollut 225. http://dx.doi.ore/10.1007/sl 1270-014-1868-2.
Labai. AL; Korosi. JB; Kurek. J; Jeziorski. A; Keller. W. (; Smol, JP. (2016). Response of
Bosmina size structure to the acidification and recovery of lakes near Sudbury, Canada.
Journal of Limnology 75: 22-29. http://dx.doi.Org/10.4081/ilimnol.2016.l 183.
Labai. AL: Kurek. J: Jeziorski. A: Smol. JP. (2015). Elevated metal concentrations inhibit
biological recovery of Cladocera in previously acidified boreal lakes. Freshw Biol 60:
347-359. http://dx.doi.org/10.Ill 1/fwb. 12496.
Lacoul. P; Freedman. B; Clair. T. (2011). Effects of acidification on aquatic biota in Atlantic
Canada. Environ Rev 19: 429-460. http://dx.doi.org/10.1139/al 1-016.
Lafrancois. BM; Carlisle. DM: Nvdick. KR; Johnson. BM; Baron. JS. (2003a). Environmental
characteristics and benthic invertebrate assemblages in Colorado mountain lakes. WestN
Am Nat 63: 137-154.
Lafrancois. BM: Nvdick. KR: Caruso. B. (2003b). Influence of nitrogen on phytoplankton
biomass and community composition in fifteen Snowy Range lakes (Wyoming, USA).
Arctic Antarct Alpine Res 35: 499-508. http://dx.doi.org/10.1657/1523-
0430(2003X)35r0499:IONOPB12.0.CO;2.
Lafrancois. BM: Nvdick. KR: Johnson. BM: Baron. JS. (2004). Cumulative effects of nutrients
and pH on the plankton of two mountain lakes. Can J Fish Aquat Sci 61: 1153-1165.
http://dx.doi.Org/10.l 139/F04-053.
Lage. MP; Reed. HE; Weihe. C: Crain. CM; Martiny. JBH. (2010). Nitrogen and phosphorus
enrichment alter the composition of ammonia-oxidizing bacteria in salt marsh sediments.
ISME J 4: 933-944. http://dx.doi.org/10.1038/ismei.201Q.10.
Laird. KR; Das. B; Kingsbury. M; Moos. MT; Pla-Rabes. S; Ahad. JME; Wiltse. B; Cumming.
BF. (2013). Paleolimnological assessment of limnological change in 10 lakes from
northwest Saskatchewan downwind of the Athabasca oils sands based on analysis of
siliceous algae and trace metals in sediment cores. Hydrobiologia 720: 55-73.
http://dx.doi.org/10.1007/slQ750-013-1623-5.
Laitha. K; Whitford. WG. (1989). The effect of water and nitrogen amendments on
photosynthesis, leaf demography, and resource-use efficiency in Larrea-tridentata, a
desert evergreen shrub. Oecologia 80: 341-348. http://dx.doi.org/10.1007/BF00379Q35.
Lamaud. E; Brunei Y; Labatut. A; Lopez. A; Fontan. J; Druilhet. A. (1994a). The Landes
experiment: Biosphere-atmosphere exchanges of ozone and aerosol particles above a pine
forest. J Geophys Res Atmos 99: 16511-16521. http://dx.doi.org/10.1029/94JD00668.
Lamaud. E; Chapuis. A; Fontan. J; Serie. E. (1994b). Measurements and parameterization of
aerosol dry deposition in a semi-arid area. Atmos Environ 28: 2461-2471.
http://dx.doi.org/10.1016/1352-2310(94^)90397-2.
June 2018
R-123
DRAFT: Do Not Cite or Quote

-------
Lamb. JB; van de Water. JA; Bourne. DG; Altier. C; Hein. MY; Fiorenza. EA; Abu. N; Join pa.
J; Harvell. CD. (2017). Seagrass ecosystems reduce exposure to bacterial pathogens of
humans, fishes, and invertebrates. Science 355: 731-733.
http://dx.doi.org/10.1126/science.aall956.
Lamers. LPM; Govers. LL; Janssen. ICJ. M; Geurts. JJM; Van Per Welle. MEW: Van Katwijk.
MM: VanDerHeide. T; Roelofs. J. anGM; Smolders. AJP. (2013). Sulfide as a soil
phytotoxin-areview. 4. http://dx.doi.org/10.3389/fpls.2013.0Q268.
Lamersdorf. NP; Beier. C; Blanck. K; Bredemeier. M; Cummins. T; Farrell. EP; Kreutzer. K;
Rasmussen. L; Ryan. M; Weis. W: Xu. Y. -J. (1998). Effect of drought experiments using
roof installations on acidification/nitrification of soils. ForEcol Manage 101: 95-109.
http://dx.doi.org/10.1016/S0378-1127(97)00128-X.
Lamsal. LN: Martin. RV; Van Donkelaar. A: Celarier. EA: Bucsela. EJ; Boersma. KF; Dirksen.
R; Luo. C; Wang. Y. (2010). Indirect validation of tropospheric nitrogen dioxide
retrieved from the OMI satellite instrument: Insight into the seasonal variation of nitrogen
oxides at northern midlatitudes. J Geophys Res 115: D05302.
http://dx.doi. org/10.1029/2009JDO13 3 51.
Lamsal. LN: Martin. RV: van Donkelaar. A: Steinbacher. M; Celarier. EA: Bucsela. E; Dunlea.
EJ; Pinto. JP. (2008). Ground-level nitrogen dioxide concentrations inferred from the
satellite-borne ozone monitoring instrument. J Geophys Res 113: D16308.
http ://dx. doi .org/10.1029/2007JD00923 5.
Lan. Z; Jenerette. GD; Zhan. S; Li. W; Zheng. S; Bai. Y. (2015). Testing the scaling effects and
mechanisms of N-induced biodiversity loss: Evidence from a decade-long grassland
experiment. J Ecol 103: 750-760. http://dx.doi.Org/10.l 111/1365-2745.12395.
Landers. DH; Eilers. JM; Brakke. DF; Overton. WS; Kellar. PE; Silverstein. ME; Schonbrod.
RD; Crowe. RE; Linthurst. RA; Omernik. JM; Teague. SA; Meier. EP. (1987). Western
lake survey phase I: Characteristics of lakes in the western United States, volume I:
Population descriptions and physico-chemical relationships. (EPA/600/3-86/054a).
Washington, DC: U.S. Environmental Protection Agency.
Landers. DH; Nahlik. AM. (2013). Final ecosystem goods and services classification system
(FEGS-CS) [EPA Report], (EPA/600/R-13/ORD-004914). Washington, DC: U.S.
Environmental Protection Agency.
https://cfpub.epa.gov/si/si public record report.cfm?dirEntryId=257922.
Landers. DH; Overton. WS; Linthurst. RA; Brakke. DF. (1988). Eastern Lake Survey: regional
estimates of lake chemistry. Environ Sci Technol 22: 128-135.
http://dx.doi.org/10.1021/esQ0167a002.
Landers. DH; Simonich. SL; Jaffe. DA; Geiser. LH; Campbell. DH; Schwindt. AR; Schreck. CB;
Kent. ML; Hafner. WD; Taylor. HE; Hageman. KJ; Usenko. S; Ackerman. LK; Schrlau.
JE; Rose. NL; Blett. TF; Erwav. MM. (2008). The fate, transport, and ecological impacts
of airborne contaminants in western national parks (USA). (EPA/600/R-07/138).
Corvallis, Oregon: U.S. Environmental Protection Agency, NHEERL, Western Ecology
Division, http://www.nature.nps.gov/air/studies/air toxics/WACAPreport.cfm.
June 2018
R-124
DRAFT: Do Not Cite or Quote

-------
Landers. DH; Simonich. SM; Jaffe. D; Geiser. L; Campbell DH; Schwindt A; Schreck. C; Kent
M; Hafner. W; Taylor. HE; Hageman, K; Usenko. S; Ackerman. L; Schrlau. J; Rose. N;
Blett, T; Erway, MM. (2010). The Western Airborne Contaminant Assessment Project
(WACAP): An interdisciplinary evaluation of the impacts of airborne contaminants in
western U.S. National Parks. Environ Sci Technol 44: 855-859.
http://dx.doi.org/10.1021/es9Q1866e.
Lane, JO; Roddam, CM; Langlois, GW; Kudela, RM. (2010). Application of solid phase
adsorption toxin tracking (SPATT) for field detection of the hydrophilic phycotoxins
domoic acid and saxitoxin in coastal California. Limnology and Oceanography: Methods
8: 645-660.
Langley, JA; McKee, KL; Cahoon, PR; Cherry, JA; Megonigal, JP. (2009). Elevated C02
stimulates marsh elevation gain, counterbalancing sea-level rise. Proc Natl Acad Sci USA
106: 6182-6186. http://dx.doi.org/10.1073/pnas.08076951Q6.
Langley, JA; Megonigal, JP. (2010). Ecosystem response to elevated C02 levels limited by
nitrogen-induced plant species shift. Nature 466: 96-99.
http://dx.doi.org/10.1038/nature09176.
Langley, JA; Megonigal, JP. (2012). Field-based radiometry to estimate tidal marsh plant growth
in response to elevated C02 and nitrogen addition. Wetlands 32: 571-578.
http://dx.doi.org/10.1007/sl3157-012-Q292-x.
Langley, JA; Mozdzer, TJ; Shepard, KA; Hagertv, SB; Megonigal, JP. (2013). Tidal marsh plant
responses to elevated C02, nitrogen fertilization, and sea level rise. Global Change Biol
19: 1495-1503. http://dx.doi.org/10.1111/gcb. 12147.
Larmola, T; Bubier, JL; Kobvlianec, C; Basiliko, N; Juutinen, S; Humphreys, E; Preston, M;
Moore, TR. (2013). Vegetation feedbacks of nutrient addition lead to a weaker carbon
sink in an ombrotrophic bog. Global Change Biol 19: 3729-3739.
http://dx.doi.org/10. Ill 1/gcb. 12328.
Larssen, T; Cosby, BJ; Lund, E; Wright, RF. (2010). Modeling future acidification and fish
populations in Norwegian surface waters. Environ Sci Technol 44: 5345-5351.
http://dx.doi.org/10.1021/esl00792m.
Latimer, JS; Charpentier, MA. (2010). Nitrogen inputs to seventy-four southern New England
estuaries: Application of a watershed nitrogen loading model. Estuar Coast Shelf Sci 89:
125-136. http://dx.doi.Org/10.1016/i.ecss.2010.06.006.
Latimer, JS; Rego, SA. (2010). Empirical relationship between eelgrass extent and predicted
watershed-derived nitrogen loading for shallow New England estuaries. Estuar Coast
Shelf Sci 90: 231-240. http://dx.doi.Org/10.1016/i.ecss.2010.09.004.
Latimer, JS; Tedesco, MA; Swanson, RL; Yarish, C; Stacev, PE; Garza, C. (2014). Long Island
sound: Prospects for the urban sea. In JS Latimer; MA Tedesco; RL Swanson; C Yarish;
PE Stacey; C Garza (Eds.), Long Island Sound: Prospects for the Urban Sea. New York,
NY: Springer. http://dx.doi.org/10.10Q7/978-l-4614-6126-5.
Lattv, EF; Canham, CD; Marks, PL. (2004). The effects of land-use history on soil properties
and nutrient dynamics in northern hardwood forests of the Adirondack Mountains.
Ecosystems 7: 193-207. http://dx.doi.org/10.1007/sl0021-003-Q157-5.
June 2018
R-125
DRAFT: Do Not Cite or Quote

-------
Laubhann, D; Sterba, H; Reinds, GJ, an; De Vries, W. im. (2009). The impact of atmospheric
deposition and climate on forest growth in European monitoring plots: An individual tree
growth model. For Ecol Manage 258: 1751-1761.
http://dx.doi.Org/10.1016/i.foreco.2008.09.050.
Laudon, H; Bishop. KH. (1999). Quantifying sources of acid neutralisation capacity depression
during spring flood episodes in Northern Sweden. Environ Pollut 105: 427-435.
http ://dx. doi .org/10.1016/S0269-7491 (99)0003 6-6.
Laudon, H; Buttle. J: Carey. SK; Mcdonnell, J: Mcguire, K; Seibert J: Shanlev, J: Soulsbv, C:
Tetzlaff. D. (2012). Cross-regional prediction of long-term trajectory of stream water
DOC response to climate change. Geophys Res Lett 39.
http://dx.doi.org/10.1029/2012GL053Q33.
Laudon. H; Norton. SA. (2010). Drivers and evolution of episodic acidification at the Bear
Brook Watershed in Maine, USA. Environ Monit Assess 171: 59-69.
http://dx.doi.org/10.1007/slQ661-010-1526-0.
Laurent. A: Fennel. K; Cai. W. eiJun; Huang. W. eiJen; Barbero. L; Wanninkhof. R. ik. (2017).
Eutrophication-induced acidification of coastal waters in the northern Gulf of Mexico:
Insights into origin and processes from a coupled physical-biogeochemical model.
Geophys Res Lett 44: 946-956. http://dx.doi.org/10.1002/2016GLQ71881.
Laursen. AE; Seitzinger. SP. (2002). The role of denitrification in nitrogen removal and carbon
mineralization in Mid-Atlantic bight sediments. Cont Shelf Res 22: 1397-1416.
Lawrence, GB; Burns, DA; Riva-Murray, K. (2016). A new look at liming as an approach to
accelerate recovery from acidic deposition effects [Review], Sci Total Environ 562: 35-
46. http://dx.doi.Org/10.1016/i.scitotenv.2016.03.176.
Lawrence, GB; David, MB; Shortle, WC. (1995). A new mechanism for calcium loss in forest-
floor soils. Nature 378: 162-165. http://dx.doi.org/10.1038/378162a0.
Lawrence, GB; David, MB; Shortle, WS; Bailey, SW; Lovett, GM. (1999). Mechanisms of base-
cation depletion by acid deposition in forest soils of the northeastern US. In SB Horsely;
RP Long (Eds.), Sugar Maple Ecology and Health: Proceedings of an International
Symposium (pp. 75-87). Warren, PA: U.S. Department of Agriculture, Forest Service,
Northeastern Research Station, http://www.treesearch.fs.fed.us/pubs/13177.
Lawrence, GB; Dukett, JE; Houck, N; Snyder, P; Capone, S. (2013). Increases in dissolved
organic carbon accelerate loss of toxic A1 in Adirondack lakes recovering from
acidification. Environ Sci Technol 47: 7095-7100. http://dx.doi.org/10.1021/es4004763.
Lawrence, GB; Hazlett, PW; Fernandez, IJ; Ouimet, R; Bailey, SW; Shortle, WC; Smith, KT;
Antidormi, MR. (2015a). Declining acidic deposition begins reversal of forest-soil
acidification in the northeastern US and eastern Canada. Environ Sci Technol 49: 13103-
13111. http://dx.doi.org/10.1021/acs.est.5b02904.
Lawrence, GB; Shortle, WC; David, MB; Smith, KT; Warbv, RAF; Lapenis, AG. (2012). Early
indications of soil recovery from acidic deposition in US red spruce forests. Soil Sci Soc
Am J 76: 1407-1417. http://dx.doi.org/10.2136/sssai2011.0415.
June 2018
R-126
DRAFT: Do Not Cite or Quote

-------
Lawrence. GB; Simonin, HA; Baldigo, BP; Roy. KM; Capone, SB. (2011). Changes in the
chemistry of acidified Adirondack streams from the early 1980s to 2008. Environ Pollut
159: 2750-2758. http://dx.doi.Org/10.1016/i.envpol.2011.05.016.
Lawrence. GB; Sullivan. TJ; Burns. DA; Bailey. SW; Cosby. BJ; Dovciak, M; Ewing, HA;
McDonnell. TC; Minocha. R; Riemann. R; Quant. J; Rice. KC; Siemion. J; Weathers.
KC. (2015b). Acidic deposition along the Appalachian Trail Corridor and its effects on
acid-sensitive terrestrial and aquatic resources: Results of the Appalachian Trail MEGA-
Transect atmospheric deposition study. (NPS/NRSS/ARD/NRR2015/996). Fort Collins,
CO: National Park Service. https://irma.nps.gov/DataStore/Reference/Profile/2223220.
Lawrence. GB; Sutherland. JW; Boylen. CW; Nierzwicki-Bauer. SW; Momen. B; Baldigo. BP;
Simonin, HA. (2007). Acid rain effects on aluminum mobilization clarified by inclusion
of strong organic acids. Environ Sci Technol 41: 93-98.
http://dx.doi.org/10.1021/es061437v.
Layer. K; Hildrew. AG; Jenkins. GB; Riede. JO; Rossiter. SJ; Townsend. CR; Woodward. G.
(2011). Long-term dynamics of a well-characterised food web: Four decades of
acidification and recovery in the broadstone stream model system. In GO Woodward
(Ed.), Ecosystems in a human-modified landscape: A European perspective (pp. 69-117).
London, UK: Academic Press. http://dx.doi.org/10.1016/B978-0-12-374794-5.000Q2-X.
Layer, K; Hildrew, AG; Woodward, G. (2013). Grazing and detritivory in 20 stream food webs
across a broad pH gradient. Oecologia 171: 459-471. http://dx.doi.org/10.1007/sQ0442-
012-2421-x.
Le Roux, X; Bouskill, NJ; Nibovet, A; Barthes, L; Dijkstra, P; Field, CB; Hungate, BA;
Lerondelle, C; Pommier, T; Tang, J; Terada, A; Tourna, M; Poly, F. (2016). Predicting
the responses of soil nitrite-oxidizers to multi-factorial global change: A trait-based
approach. FMICB 7: 628. http://dx.doi.org/10.3389/fmicb.2016.00628.
Leavitt, PR; Findlav, PL; Hall, RI; Smol, JP. (1999). Algal responses to dissolved organic
carbon loss and pH decline during whole-lake acidification: Evidence from
paleolimnology. Limnology 44: 757-773.
http://dx.doi.Org/10.4319/lo.1999.44.3 part 2.0757.
LeBauer, DS; Treseder, KK. (2008). Nitrogen limitation of net primary productivity in terrestrial
ecosystems is globally distributed. Ecology 89: 371-379. http://dx.doi.org/10.1890/06-
2057.1.
Leblanc, GA. (1995). Trophic-level differences in the bioconcentration of chemicals:
Implications in assessing environmental biomagnification. Environ Sci Technol 29: 154-
160.
Lebo, ME; Paerl, HW; Peierls, BL. (2012). Evaluation of progress in achieving TMDL mandated
nitrogen reductions in the Neuse River Basin, North Carolina. Environ Manage 49: 253-
266. http://dx.doi.org/10.1007/s00267-011 -9774-5.
Ledesma, JLJ; Futter, MN; Laudon, H; Evans, CD; Kohler, SJ. (2016). Boreal forest riparian
zones regulate stream sulfate and dissolved organic carbon. Sci Total Environ 560: 110-
122. http://dx.doi.Org/10.1016/i.scitotenv.2016.03.230.
June 2018
R-127
DRAFT: Do Not Cite or Quote

-------
Leduc. AOH. C; Roh. E; Brown. GE. (2009). Effects of acid rainfall on juvenile Atlantic salmon
(Salmo salar) antipredator behaviour: Loss of chemical alarm function and potential
survival consequences during predation. Mar Freshwat Res 60: 1223-1230.
http://dx.doi.org/10.1071/MFQ8323.
Lee. C: Martin. RV; van Donkelaar. A: Lee. H: Dickerson. RR; Hains. JC: Krotkov. N: Richter.
A: Vinnikov. K: Schwab. JJ. (2011). S02 emissions and lifetimes: Estimates from
inverse modeling using in situ and global, space-based (SCIAMACHY and OMI)
observations. J Geophys Res 116: D06304. http://dx.doi.org/10.1029/2010JDQ14758.
Lee. HM: Paulot. F; Henze. DK; Travis. K; Jacob. DJ: Pardo. LH: Schichtel. BA. (2016).
Sources of nitrogen deposition in Federal Class I areas in the US. Atmos Chem Phys 16:
525-540. http://dx.doi.org/10.5194/acp-16-525-2016.
Lee. JA; Studholme. CJ. (1992). Responses of Sphagnum species to polluted environments. In
JW Bates; AM Farmer (Eds.), Bryophytes and Lichens in a Changing Environment (pp.
314-332). Oxford, England: Clarendon Press.
Lee. S: Yoo. S. (2016). Interannual variability of the phytoplankton community by the changes
in vertical mixing and atmospheric deposition in the Ulleung Basin, East Sea: A
modelling study. Ecol Modell 322: 31-47.
http://dx.doi.Org/10.1016/i.ecolmodel.2015.l 1.012.
Lee. T; Yu. X. -Y; Ayres. B; Kreidenweis. SM; Malm. WC: Collett. JL. (2008). Observations of
fine and coarse particle nitrate at several rural locations in the United States. Atmos
Environ 42: 2720-2732. http://dx.doi.Org/10.1016/i.atmosenv.2007.05.016.
Lee. TP: Tioelker. MG: Ellsworth. DS: Reich. PB. (2001). Leaf gas exchange responses of 13
prairie grassland species to elevated C02 and increased nitrogen supply. New Phytol
150: 405-418. http://dx.doi.org/10.1046/i. 1469-8137.2001.00095.x.
Lee, WY; Iannucci-Berger, WA; Eitzer, BP; White, JC; Mattina, MI. (2003). Plant uptake and
translocation of air-borne chlordane and comparison with the soil-to-plant route.
Chemosphere 53: 111-121. http://dx.doi.org/10.1016/S0045-6535(03N)00353-9.
Lefcheck, JS: Wilcox, PJ; Murphy, RR: Marion, SR; Orth, RJ. (In Press) Multiple stressors
threaten the imperiled coastal foundation species eelgrass (Zostera marina) in Chesapeake
Bay, USA. Global Change Biol, http://dx.doi.org/10. Ill 1/gcb. 13623.
Lefebvre, SC: Benner, I: Stillman, JH; Parker, AE; Prake, MK; Rossignol, PE; Okimura, KM:
Komada, T; Carpenter, EJ. (2012). Nitrogen source and pCO(2) synergistically affect
carbon allocation, growth and morphology of the coccolithophore Emiliania huxleyi:
Potential implications of ocean acidification for the carbon cycle. Global Change Biol 18:
493-503. http://dx.doi.org/10.1111/i. 1365-2486.2011.02575.x.
Lefer, BL; Talbot, RW. (2001). Summertime measurements of aerosol nitrate and ammonium at
a northeastern U.S. site. J Geophys Res 106: 20365-20378.
http://dx.doi.org/10.1029/2000JP90Q693.
Legg, BJ: Powell, FA. (1979). Spore dispersal in a barley crop: A mathematical model. Agr
Meteorol 20: 47-67. http://dx.doi.org/10.1016/0002-1571 (79^)90050-5.
June 2018
R-128
DRAFT: Do Not Cite or Quote

-------
Lehman. PW; Bover. G; Satchwell. M; Waller. S. (2008). The influence of environmental
conditions on the seasonal variation of Microcystis cell density and microcystins
concentration in San Francisco Estuary. Hydrobiologia 600: 187-204.
http://dx.doi.org/10.1007/sl0750-0Q7-9231-x.
Lehman. PW: Kendall. C: Guerin. MA: Young. MB: Silva. SR. SR; Bover. GL; Teh. SJ. (2015).
Characterization of the microcystis bloom and its nitrogen supply in San Francisco
Estuary using stable isotopes. Estuaries Coasts 38: 165-178.
http://dx.doi.org/10.1007/sl2237-014-9811-8.
Lehmann. J: Solomon. D: Zhao. FJ: McGrath. SP. (2008). Atmospheric S02 emissions since the
late 1800s change organic sulfur forms in humic substance extracts of soils. Environ Sci
Technol 42: 3550-3555. http://dx.doi.org/10.1021/es702315g.
Lehnherr. I; St. Louis. VL; Hintelmann. H: Kirk. JL. (2011). Methylation of inorganic mercury
in polar marine waters. Nat Geosci 4: 298-302. http://dx.doi.org/10.1038/NGEQ1134.
Leivestad. H. (1982). Physiological effects of acid stress on fish. In TA Haines; E Johnson
(Eds.), Acid Rain/Fisheries. Bethesda, MD: Northeastern Division American Fisheries
Society, http://fisheries.org/shop/x53007xm.
Lenart-Boron, A; Wolny-Koladka, K. (2015). Heavy metal concentration and the occurrence of
selected microorganisms in soils of a steelworks area in Poland. Plant Soil Environ 61:
273-278. http://dx.doi.org/10.17221/217/2015-PSE.
Lendzian. KJ; Kerstiens. G. (1988). Interactions between plant cuticles and gaseous air
pollutants. Aspect Appl Biol 17: 97-104.
Lento. J: Dillon. PJ; Somers. KM. (2012). Evaluating long-term trends in littoral benthic
macroinvertebrate communities of lakes recovering from acid deposition. Environ Monit
Assess 184: 7175-7187. http://dx.doi.org/10.1007/slQ661-011-2489-5.
Lento. J: Dillon. PJ: Somers. KM: Reid. RA. (2008). Changes in littoral benthic
macroinvertebrate communities in relation to water chemistry in 17 Precambrian Shield
lakes. Can J Fish Aquat Sci 65: 906-918. http://dx.doi.Org/10.l 139/F08-033.
Leonardi. S: Gentilesca. T; Guerrieri. R; Ripullone. F; Magnani. F; Mencuccini. M; Noiie. TV:
Borghetti. M. (2012). Assessing the effects of nitrogen deposition and climate on carbon
isotope discrimination and intrinsic water-use efficiency of angiosperm and conifer trees
under rising C02 conditions. Global Change Biol 18: 2925-2944.
http://dx.doi.org/10.111 l/i,1365-2486.2012.02757.x.
Lepori. F; Robin. J. (2014). Nitrogen limitation of the phytobenthos in Alpine lakes: Results
from nutrient-diffusing substrata. FreshwBiol 59: 1633-1645.
http://dx.doi.org/10. Ill 1/fwb. 12370.
Lesack. LFW: Melack. JM. (1996). Mass balance of major solutes in a rainforest catchment in
the central Amazon: Implications for nutrient budgets in tropical rainforests.
Biogeochemistry 32: 115-142. http://dx.doi.org/10.1007/bfD000Q355.
Lesser, MR; Fridley, JD. (2016). Global change at the landscape level: Relating regional and
landscape-scale drivers of historical climate trends in the Southern Appalachians. Int J
Climatol 36: 1197-1209. http://dx.doi.org/10.10Q2/ioc.4413.
June 2018
R-129
DRAFT: Do Not Cite or Quote

-------
Leveque. T; Capowiez. Y; Schreck. E. va; Xiong. T; Foucault Y; Dumat C. (2014). Earthworm
bioturbation influences the phytoavailability of metals released by particles in cultivated
soils. Environ Pollut 191: 199-206. http://dx.doi.Org/10.1016/i.envpol.2014.04.005.
Levi a. DF; Frost EE. (2003). A review and evaluation of stemflow literature in the hydrologic
and biogeochemical cycles of forested and agricultural ecosystems. J Hydrol 274: 1-29.
http://dx.doi. org/10.1016/S0022-1694(02^)003 99-2.
Levin. LA: Ekau. W: Goodav. AJ; Jorissen. F; Middelburg. JJ; Nacivi. SWA: Neira. C: Rabalais.
NN; Zhang. J. (2009). Effects of natural and human-induced hypoxia on coastal benthos.
Biogeosciences 6: 2063-2098. http://dx.doi.org/10.5194/bg-6-2063-2009.
Levy. PE; Wendler. R; Van Oiien. M; Cannell. MGR.; Millard. P. (2005). The effect of nitrogen
enrichment on the carbon sink in coniferous forests: Uncertainty and sensitivity analyses
of three ecosystem models. Water, Air, and Soil Pollution: Focus 4: 67-74.
http://dx.doi.org/10.1007/sll267-005-3Q15-v.
Lewis. WM; Wurtsbaugh. WA. (2008). Control of lacustrine phytoplankton by nutrients: Erosion
of the phosphorus paradigm [Review], Int Rev Hydrobiol 93: 446-465.
http://dx.doi.org/10.1002/iroh.20Q811065.
Lewis, WM; Wurtsbaugh, WA; Paerl, HW. (2011). Rationale for control of anthropogenic
nitrogen and phosphorus to reduce eutrophication of inland waters. Environ Sci Technol
45: 10300-10305. http://dx.doi.org/10.1021/es202401p.
Li, C: Joiner, J: Krotkov, NA; Bhartia, PK. (2013). A fast and sensitive new satellite S02
retrieval algorithm based on principal component analysis: Application to the ozone
monitoring instrument. Geophys Res Lett 40: 6314-6318.
http://dx.doi.org/10.1002/2013GLQ58134.
Li, H: McNultv, SG. (2007). Uncertainty analysis on simple mass balance model to calculate
critical loads for soil acidity. Environ Pollut 149: 315-326.
http://dx.doi.Org/10.1016/i.envpol.2007.05.014.
Li, H, ui; Xu, Z; Yang, S: Li, X: Top, E; Wang, R; Zhang, Y; Cai, J: Yao, F, ei; Han, X: Jiang,
Y. (2016a). Responses of Soil Bacterial Communities to Nitrogen Deposition and
Precipitation Increment Are Closely Linked with Aboveground Community Variation.
Microb Ecol 71: 974-989. http://dx.doi.org/10.1007/s00248-016-073Q-z.
Li, H, ui: Yang, S: Xu, Z; Yan, O; Li, X: van Nostrand, J: He, Z; Yao, F, ei: Han, X: Zhou, J:
Deng, Y, e: Jiang, Y. (2017a). Responses of soil microbial functional genes to global
changes are indirectly influenced by aboveground plant biomass variation. Soil Biol
Biochem 104: 18-29. http://dx.doi.Org/10.1016/i.soilbio.2016.10.009.
Li. M: Lee. YJ: Testa. JM: Li. Y. un: Ni. W: Kemp. WM: Pi Toro. DM. (2016b). What drives
interannual variability of hypoxia in Chesapeake Bay: Climate forcing versus nutrient
loading? Geophys Res Lett 43: 2127-2134. http://dx.doi.org/10.1002/2015GLQ67334.
Li. S: Bush. RT; Mao. R; Xiong. L; Ye. C. (2017b). Extreme drought causes distinct water
acidification and eutrophi cation in the Lower Lakes (Lakes Alexandrina and Albert),
Australia. J Hydrol 544: 133-146. http://dx.doi.Org/10.1016/i.ihydrol.2016.l 1.015.
June 2018
R-130
DRAFT: Do Not Cite or Quote

-------
Li. S; Mendelssohn. IA; Chen. H; Orem, WH. (2009). Does sulphate enrichment promote the
expansion of Typha domingensis (cattail) in the Florida Everglades?(l). Freshw Biol 54:
1909-1923. http://dx.doi.org/10.1111/j. 1365-2427.2009.02242.X.
Li. W: Jin. C: Guan. D; Wang. 0; Wang. A: Yuan. F; Wu. J. (2015). The effects of simulated
nitrogen deposition on plant root traits: A meta-analysis. Soil Biol Biochem 82: 112-118.
http://dx.doi.Org/10.1016/i.soilbio.2015.01.001.
Li. W: Wu. J: Bai. E; Guan. D; Wang. A: Yuan. F; Wang. S: Jin. C. (2016c). Response of
terrestrial nitrogen dynamics to snow cover change: A meta-analysis of experimental
manipulation. Soil Biol Biochem 100: 51-58.
http://dx.doi.Org/10.1016/i.soilbio.2016.05.018.
Li. X: Weller. DE; Gallegos. CL; Jordan. TE; Kim. H. -C. (2007). Effects of watershed and
estuarine characteristics on the abundance of submerged aquatic vegetation in
Chesapeake Bay subestuaries. Estuaries Coasts 30: 840854.
http://dx.doi. org/10.1007/BF02841338.
Li. XY; Meixner. T; Sickman. JO: Miller. AE; Schimel. JP; Melack. JM. (2006). Decadal-scale
dynamics of water, carbon and nitrogen in a California chaparral ecosystem: DAYCENT
modeling results. Biogeochemistry 77: 217-245. http://dx.doi.org/10.1007/slQ533-005-
1391-z.
Li. Y; Schichtel. BA; Walker. JT; Schwede. DB; Chen. X: Lehmann. CM: Puchalski. MA: Gay.
DA: Collett. JL. Jr. (2016d). Increasing importance of deposition of reduced nitrogen in
the United States. Proc Natl Acad Sci USA 113: 5874-5879.
http://dx.doi.org/10.1073/pnas.1525736113.
Li. YY; Lti. X. -T; Wang. Z. -W: Zhou. C: Han. X. -G. (2014). Linking relative growth rates to
biomass allocation: The responses of the grass Leymus chinensis to nitrogen addition.
Phyton (Buenos Aires) 83: 283-289.
Liang. C: Gutknecht. JLM; Balser. TC. (2015). Microbial lipid and amino sugar responses to
long-term simulated global environmental changes in a California annual grassland.
FMICB 6: 385. http://dx.doi.org/10.3389/fmicb.2015.00385.
Liao, J: Froyd, KD; Murphy, DM; Keutsch, FN: Yu, G; Wennberg, PO; St. Clair, JM; Crounse,
JD; Wisthaler, A: Mikoviny, T; Jimenez, JL: Campuzano-Jost, P; Day, DA: Hu, W:
Ryerson, TB; Pollack, IB: Peischl, J: Anderson, BE: Ziemba, LP: Blake, PR: Meinardi,
S: Pi skin, G. (2015). Airborne measurements of organosulfates over the continental US. J
Geophys Res Atmos 120: 2990-3005. http://dx.doi.org/10.1002/2014JPQ22378.
Lichter, J: Barron, SH: Bevacqua, CE: Finzi, AC: Irving, KE: Stemmler, EA: Schlesinger, WH.
(2005). Soil carbon sequestration and turnover in a pine forest after six years of
atmospheric C02 enrichment. Ecology 86: 1835-1847. http://dx.doi.org/10.1890/Q4-
1205.
Lieb, AM: Parrouzet-Nardi, A: Bowman, WP. (2011). Nitrogen deposition decreases acid
buffering capacity of alpine soils in the southern Rocky Mountains. Geoderma 164: 220-
224. http://dx.doi.Org/10.1016/i.geoderma.2011.06.013.
June 2018
R-131
DRAFT: Do Not Cite or Quote

-------
Liebich. T; McCormick. SD; Kircheis. D; Johnson. K; Regal. R; Hrabik. T. (2011). Water
chemistry and its effects on the physiology and survival of Atlantic salmon Salmo salar
smolts. J Fish Biol 79: 502-519. http://dx.doi.org/10.1111/j. 1095-8649.2011.03046.x.
Lien. L; Raddum. GG: Fiellheim. A: Henriksen. A. (1996). A critical limit for acid neutralizing
capacity in Norwegian surface waters, based on new analyses of fish and invertebrate
responses. Sci Total Environ 177: 173-193. http://dx.doi.org/10.1016/0Q48-
9697(95)04894-4.
Liess. A: Drakare. S: Kahlert. M. (2009). Atmospheric nitrogen-deposition may intensify
phosphorus limitation of shallow epilithic periphyton in unproductive lakes. Freshw Biol
54: 1759-1773. http://dx.doi.Org/10.llll/i.1365-2427.2009.02222.x.
Liggio. J: Li. SM; Vlasenko. A: Stroud. C: Makar. P. (2011). Depression of ammonia uptake to
sulfuric acid aerosols by competing uptake of ambient organic gases. Environ Sci
Technol 45: 2790-2796. http://dx.doi.org/10.1021/esl03801g.
Likar. M; Regvar. M. (2009). Application of temporal temperature gradient gel electrophoresis
for characterisation of fungal endophyte communities of Salix caprea L. in a heavy metal
polluted soil. Sci Total Environ 407: 6179-6187.
http://dx.doi.Org/10.1016/i.scitotenv.2009.08.045.
Likens. GE; Butler. TJ; Buso. DC. (2001). Long- and short-term changes in sulfate deposition:
Effects of the 1990 Clean Air Act Amendments. Biogeochemistry 52: 1-11.
http://dx.doi.Org/10.1023/A:1026563400336.
Likens, GE; Driscoll, CT; Buso, DC. (1996). Long-term effects of acid rain: response and
recovery of a forest ecosystem. Science 272: 244-246.
http://dx.doi.Org/10.l 126/science.272.5259.244.
Lilleskov, EA. (1999) Decline of above-and belowground ectomycorrhizal fungal diversity over
an atmospheric nitrogen deposition gradient near Kenai, Alaska. (Doctoral Dissertation).
Cornell University, Ithaca, NY. Retrieved from
https://newcatalog.library.cornell.edu/catalog/3449564
Lilleskov, EA: Fahev, TJ: Horton, TR; Lovett, GM. (2002). Belowground ectomycorrhizal
fungal community change over a nitrogen deposition gradient in Alaska. Ecology 83:
104-115. http://dx.doi.org/10.1890/0012-9658(2002)083r0104:BEFCC012.0.CQ:2.
Lilleskov, EA: Fahev, TJ: Lovett, GM. (2001). Ectomycorrhizal fungal aboveground community
change over an atmospheric nitrogen deposition gradient. Ecol Appl 11: 397-410.
http://dx.doi.org/10.1890/1051-0761(2001)01 ir0397:EFACCQ12.0.CQ:2.
Lilleskov, EA: Hobbie, EA: Horton, TR. (2011). Conservation of ectomycorrhizal fungi:
Exploring the linkages between functional and taxonomic responses to anthropogenic N
deposition. Fungal Ecology 4: 174-183. http://dx.doi.Org/10.1016/i.funeco.2010.09.008.
Lilleskov, EA: Wargo, PM; Vogt, KA; Vogt, DJ. (2008). Mycorrhizal fungal community
relationship to root nitrogen concentration over a regional atmospheric nitrogen
deposition gradient in the northeastern USA. Can J For Res 38: 1260-1266.
http://dx.doi.org/10.1139/X07-211.
June 2018
R-132
DRAFT: Do Not Cite or Quote

-------
Lima-Fernandes, E, va; Fernandes, I; Pereira, A. na; Geraldes, P; Cassio, F; Pascoal, C. (2015).
Eutrophication modulates plant-litter diversity effects on litter decomposition in streams.
Freshw Sci 34: 31-41. http://dx.doi.ore/10.1086/679223.
Limpens, J: Berendse, F; Klees, H. (2003 a). N deposition affects N availability in interstitial
water, growth of Sphagnum and invasion of vascular plants in bog vegetation. New
Phytol 157: 339-347. http://dx.doi.Org/10.1046/i.1469-8137.2003.00667.x.
Li m pens. J: Gran at h. G: Gunnarsson. U: Aerts. R: Bavlev. S: Bragazza, L: Bubier. J: Buttler. A:
van Den Berg. LJL: Francez. AJ: Gerdol. R: Grosvernier. P; Heiimans. MMP. D:
Hoosbeek, MR: Hotes, S: Hornets. M; Leith, I; Mitchell EAD: Moore. T; Nilsson, MB:
Nordbakken. JF; Rochefort. L: Rydin. H: Sheppard. LJ; Thormann. M; Wiedemann.
MM; Williams, BL; Xu, B. (2011). Climatic modifiers of the response to nitrogen
deposition in peat-forming Sphagnum mosses: a meta-analysis. New Phytol 191: 496-
507. http://dx.doi.org/10.1111/i. 1469-8137.2011,03680.x.
Limpens. J: Ravmakers. JTA. G: Baar. J: Berendse. F; Zijlstra. JD. (2003b). The interaction
between epiphytic algae, a parasitic fungus and Sphagnum as affected by N and P. Oikos
103: 59-68. http://dx.doi.Org/10.1034/i.1600-0706.2003.12580.x.
Lin. H: He. Z: Hao. J: Tian. K. ai: Jia. X: Kong. X: Akbar. S: Bei. Z: Tian. X. (2017). Effect of N
addition on home-field advantage of litter decomposition in subtropical forests. For Ecol
Manage 398: 216-225. http://dx.doi.Org/10.1016/i.foreco.2017.05.015.
Linder. G: Brumbaugh. W: Neitlich. P; Little. E. (2013). Atmospheric deposition and critical
loads for nitrogen and metals in arctic Alaska: Review and current status. 02: 76-99.
http://dx.doi.org/10.4236/oiap.2013.24010.
Linker. LC: Dennis. R; Shenk. GW: Batiuk. RA; Grimm. J: Wang. P. (2013). Computing
atmospheric nutrient loads to the Chesapeake bay watershed and tidal waters. J Am Water
Resour Assoc 49: 1025-1041. http://dx.doi.org/10. Ill 1/iawr. 12112.
Lisa. JA; Song. B; Tobias. CR; Duernberger. KA. (2014). Impacts of freshwater flushing on
anammox community structure and activities in the New River Estuary, USA. Aquat
Microb Ecol 72: 17-31. http://dx.doi.org/10.3354/ame01682.
Lisa, JA; Song, B; Tobias, CR: Hines, DE. (2015). Genetic and biogeochemical investigation of
sedimentary nitrogen cycling communities responding to tidal and seasonal dynamics in
Cape Fear River Estuary. Estuar Coast Shelf Sci 167: A313-A323.
http://dx.doi.Org/10.1016/i.ecss.2015.09.008.
LISS (Long Island Sound Study). (2017). Long Island Sound study (LISS). Available online at
http://longislandsoundstudv.net/indicator/water-qualitv-index/ (accessed August 23,
2017).
Litton, CM: Raich, JW: Ryan, MG. (2007). Carbon allocation in forest ecosystems. Global
Change Biol 13: 2089-2109. http://dx.doi.org/10.111 l/i.l365-2486.2007.01420.x.
Liu, J: Wu, N: Wang, H; Sun, J: Peng, B; Jiang, P; Bai, E. (2016a). Nitrogen addition affects
chemical compositions of plant tissues, litter and soil organic matter. Ecology 97: 1796-
1806. http://dx.doi.Org/10.1890/15-1683.l.
June 2018
R-133
DRAFT: Do Not Cite or Quote

-------
Liu. K; Crowley. D. (2009). Nitrogen deposition effects on carbon storage and fungal:bacterial
ratios in coastal sage scrub soils of Southern California. J Environ Qual 38: 2267-2272.
http://dx.doi.ore/10.2134/ieq2008.0387.
Liu. L; Greaver. TL. (2009). A review of nitrogen enrichment effects on three biogenic GHGs:
The C02 sink may be largely offset by stimulated N20 and CH4 emission [Review],
Ecol Lett 12: 1103-1117. http://dx.doi.Org/10.llll/i.1461-0248.2009.01351.x.
Liu. L; Greaver. TL. (2010). A global perspective on belowground carbon dynamics under
nitrogen enrichment. Ecol Lett 13: 819-828. http://dx.doi.Org/10.l 11 l/j.1461-
0248.2010.01482.x.
Liu. L. in: Monaco. TA; Sun. F; Liu. W. ei; Gan. Y; Sun. G. (2017a). Altered precipitation
patterns and simulated nitrogen deposition effects on phenology of common plant species
in a Tibetan Plateau alpine meadow. Agr Forest Meteorol 236: 36-47.
http://dx.doi.Org/10.1016/i.agrformet.2017.01.010.
Liu. L; Wang. X: Laieunesse. MJ; Miao. G: Piao. S: Wan. S: Wu. Y; Wang. Z; Yang. S: Li. P;
Deng. M. (2016b). A cross-biome synthesis of soil respiration and its determinants under
simulated precipitation changes. Global Change Biol 22: 1394-1405.
http://dx.doi.org/10. Ill 1/gcb. 13156.
Liu. Y; Evans. MA: Scavia. D. (2010). Gulf of Mexico hypoxia: Exploring increasing sensitivity
to nitrogen loads. Environ Sci Technol 44: 5836-5841.
http://dx.doi.org/10.1021/es903521n.
Liu, Y; Miao, R; Chen, A; Miao, Y; Liu, Y; Wu, X. (2017b). Effects of nitrogen addition and
mowing on reproductive phenology of three early-flowering forb species in a Tibetan
alpine meadow. Ecol Eng 99: 119-125. http://dx.doi.Org/10.1016/i.ecoleng.2016.l 1.033.
Liu, Y, un; Chen, T; Song, S: Li, C. (2015). Effects of nitrogenous nutrition on growth and
nitrogen assimilation enzymes of dinoflagellate Akashiwo sanguinea. Harmful Algae 50:
99-106. http://dx.doi.Org/10.1016/i.hal.2015.10.005.
Lo, BP: Elphick, JR; Bailey, HC: Baker, JA; Kennedy, CJ. (2015). The effect of sulfate on
selenate bioaccumulation in two freshwater primary producers: A duckweed (Lemna
minor) and a green alga (Pseudokirchneriella subcapitata). Environ Toxicol Chem 34:
2841-2845. http://dx.doi.org/10.1002/etc.3127.
Locke, A: Sprules, WG. (1994). Effects of lake acidification and recovery on the stability of
zooplankton food webs. Ecology 75: 498-506. http://dx.doi.org/10.2307/1939553.
Lockwood, AL; Shepson, PB; Fiddler, MN: Alaghmand, M. (2010). Isoprene nitrates:
Preparation, separation, identification, yields, and atmospheric chemistry. Atmos Chem
Phys 10: 6169-6178. http://dx.doi.org/10.5194/acp-10-6169-2010.
Lofgren, S: Gustafsson, J, onP; Bringmark, L. (2010). Decreasing DOC trends in soil solution
along the hillslopes at two IM sites in southern Sweden - Geochemical modeling of
organic matter solubility during acidification recovery. Sci Total Environ 409: 201-210.
http://dx.doi.Org/10.1016/i.scitotenv.2010.09.023.
June 2018
R-134
DRAFT: Do Not Cite or Quote

-------
Logofet, DO: Alexandrov. GA. (1984). Modelling of matter cycle in a mesotrophic bog
ecosystem II. Dynamic model and ecological succession. Ecol Modell 21: 259-276.
http://dx.doi. org/10.1016/03 04-3 800(84)90063 -2.
Lohman. K: Jones. JR: Bavsinger-Daniel. C. (1991). Experimental evidence for nitrogen
limitation in a northern Ozark stream. J North Am Benthol Soc 10: 14-23.
http://dx.doi.org/10.2307/1467760.
Lohman. K: Priscu, JC. (1992). Physiological indicators of nutrient deficiency in cladophora
(chlorophyta) in the Clark Fork of the Columbia River, Montana. J Phycol 28: 443-448.
http://dx.doi.org/10.1111/i .0022-3646.1992.00443.x.
Long. RP: Horslev. SB: Hallett. RA: Bailey. SW. (2009). Sugar maple growth in relation to
nutrition and stress in the northeastern United States. Ecol Appl 19: 1454-1466.
http://dx.doi.Org/10.1890/08-1535.l.
Lopez, CB: Jewett, EB: Dortch, Q: Walton, BT: Hudnell, HK. (2008). Scientific assessment of
freshwater harmful algal blooms. Washington, DC: Interagency Working Group on
Harmful Algal Blooms, Hypoxia, and Human Health of the Joint Subcommittee on
Ocean Science and Technology.
https://obamawhitehouse.archives.gov/sites/default/files/microsites/ostp/frshh2o0708.pdf.
Loreau, M: de Mazancourt, C. (2013). Biodiversity and ecosystem stability: A synthesis of
underlying mechanisms. Ecol Lett 16: 106-115. http://dx.doi.Org/10.l 111/ele. 12073.
Lorenz, R: Murphy, CE, Jr. (1989). Dry deposition of particles to a pine plantation. Boundary
Layer Meteorol 46: 355-366. http://dx.doi.org/10.1007/BF0Q172241.
Loubet, B: Milford, C: Sutton, MA: Cellier, P. (2001). Investigation of the interaction between
sources and sinks of atmospheric ammonia in an upland landscape using a simplified
dispersion-exchange model. J Geophys Res Atmos 106: 24183-24195.
http://dx.doi. org/10.1029/2001JD90023 8.
Lovelock, CE: Ball, MC: Martin, KC: Feller, IC. (2009). Nutrient enrichment increases mortality
of mangroves. PLoS ONE 4: e5600. http://dx.doi.org/10.1371/iournal.pone.000560Q.
Lovett, GM. (2013). Critical issues for critical loads [Comment], Proc Natl Acad Sci USA 110:
808-809. http://dx.doi.org/10.1073/pnas.121900711Q.
Lovett, GM. (1992). Atmospheric deposition and canopy interactions of nitrogen. In DW
Johnson; SE Lindberg (Eds.), Atmospheric Deposition and Forest Nutrient Cycling: A
Synthesis of the Integrated Forest Study (v. 91 ed., pp. 152-166). New York, NY:
Springer-Verlag, Inc. http://dx.doi.org/10.1007/978-l-4612-2806-6.
Lovett, GM: Arthur, MA: Weathers, KC: Fitzhugh, RD: Templer, PH. (2013). Nitrogen addition
increases carbon storage in soils, but not in trees, in an Eastern U.S. deciduous forest.
Ecosystems 16: 980-1001. http://dx.doi.org/10.1007/slQ021-013-9662-3.
Lovett, GM: Goodale, CL. (2011). A new conceptual model of nitrogen saturation based on
experimental nitrogen addition to an oak forest. Ecosystems 14: 615-631.
http://dx.doi.org/10.1007/slQ021-011-9432-z.
June 2018
R-135
DRAFT: Do Not Cite or Quote

-------
Lovett GM; Tear. TH; Evers, DC; Findlav, SEG; Cosby. BJ; Dunscomb, JK; Driscoll, CT;
Weathers. KC. (2009). Effects of air pollution on ecosystems and biological diversity in
the eastern United States [Review], AnnN Y Acad Sci 1162: 99-135.
http://dx.doi.org/10.1111/U749-6632.2009.04153.X.
Lozanovska. I: Kuzyakov. Y; Krohn. J: Parvin. S: Dorodnikov. M. (2016). Effects of nitrate and
sulfate on greenhouse gas emission potentials from microform-derived peats of a boreal
peatland: A C-13 tracer study. Soil Biol Biochem 100: 182-191.
http://dx.doi.Org/10.1016/i.soilbio.2016.06.018.
Lu. C: Tian. H. (2013). Net greenhouse gas balance in response to nitrogen enrichment:
Perspectives from a coupled biogeochemical model. Global Change Biol 19: 571-588.
http://dx.doi.org/10. Ill 1/gcb. 12049.
Lu. C: Tian. H; Liu. M; Ren. W. ei; Xu. X: Chen. G: Zhang. C. (2012). Effect of nitrogen
deposition on China's terrestrial carbon uptake in the context of multifactor
environmental changes. Ecol Appl 22: 53-75. http://dx.doi.Org/10.1890/10-1685.l.
Lu. M; Yang. Y; Luo. Y; Fang. C: Zhou. X: Chen. J: Yang. X: Li. B. (201 la). Responses of
ecosystem nitrogen cycle to nitrogen addition: A meta-analysis. New Phytol 189: 1040-
1050. http://dx.doi.org/10.1111/i. 1469-8137.2010.03563,x.
Lu. M; Zhou. X: Luo. Y; Yang. Y; Fang. C: Chen. J: Li. B. (201 lb). Minor stimulation of soil
carbon storage by nitrogen addition: A meta-analysis. Agric Ecosyst Environ 140: 234-
244. http://dx.doi.Org/10.1016/i.agee.2010.12.010.
Lu, M; Zhou, X; Yang, Q; Li, H; Luo, Y; Fang, C; Chen, J; Yang, X; Li, B. (2013). Responses of
ecosystem carbon cycle to experimental warming: A meta-analysis. Ecology 94: 726-738.
http://dx.doi.Org/10.1890/12-0279.l.
Lu, X: Jiang, H: Liu, J: Zhang, X: Jin, J: Zhu, Q: Zhang, Z; Peng, C. (2016). Simulated effects of
nitrogen saturation on the global carbon budget using the IBIS model. Sci Rep 6: 39173.
http://dx.doi.org/10.1038/srep39173.
Lu, X: Mo, J: Gilliam, FS: Yu, G: Zhang, W: Fang, Y; Huang, J. (201 lc). Effects of
experimental nitrogen additions on plant diversity in tropical forests of contrasting
disturbance regimes in southern China. Environ Pollut 159: 2228-2235.
http://dx.doi.Org/10.1016/i.envpol.2010.10.037.
Lu, X: Mo, J: Gilliam, FS: Zhou, G: Fang, Y. (2010). Effects of experimental nitrogen additions
on plant diversity in an old-growth tropical forest. Global Change Biol 16: 2688-2700.
http://dx.doi.org/10.111 l/i.l365-2486.2010.02174.x.
Lu, XT, ao: Reed, S: Hou, SL, i; Hu, Y, anYu; Wei, H, aiWei; Lu, F, uMei; Cui, O; Han, XG,
uo. (2017). Temporal variability of foliar nutrients: responses to nitrogen deposition and
prescribed fire in a temperate steppe. Biogeochemistry 133: 295-305.
http://dx.doi.org/10.1007/slQ533-017-0333-x.
Lucas, RW: Holmstrom, H; Lamas, T. (2014). Intensive forest harvesting and pools of base
cations in forest ecosystems: A modeling study using the Heureka decision support
system. For Ecol Manage 325: 26-36. http://dx.doi.Org/10.1016/i.foreco.2014.03.053.
June 2018
R-136
DRAFT: Do Not Cite or Quote

-------
Lucas. RW; Klaminder. J: Futter. MN; Bishop. KH; Egnell. G; Laudon. H; Hogberg. P. (2011).
A meta-analysis of the effects of nitrogen additions on base cations: Implications for
plants, soils, and streams. For Ecol Manage 262: 95-104.
http://dx.doi.Org/10.1016/i.foreco.2011.03.018.
Lucash. MS: Yanai. RD; Blum. JD; Park. BB. (2012). Foliar nutrient concentrations related to
soil sources across a range of sites in the northeastern United States. Soil Sci Soc Am J
76: 674-683. http://dx.doi.org/10.2136/sssai2011.016Q.
Luo. Y; Guo. W: Yuan. Y; Liu. J: Du. N: Wang. R. (2014). Increased nitrogen deposition
alleviated the competitive effects of the introduced invasive plant Robinia pseudoacacia
on the native tree Quercus acutissima. Plant Soil 385: 63-75.
http://dx.doi.org/10.1007/sl 1104-014-2227-1.
Lusardi. RA; Bogan. MT; Movie. PB; Dahlgren. RA. (2016). Environment shapes invertebrate
assemblage structure differences between volcanic spring-fed and runoff rivers in
northern California. Freshw Sci 35: 1010-1022. http://dx.doi.org/10.1086/687114.
Lydersen. E; Larssen. T; Field. E. (2004). The influence of total organic carbon (TOC) on the
relationship between acid neutralizing capacity (ANC) and fish status in Norwegian
lakes. Sci Total Environ 326: 63-69. http://dx.doi.org/10.1016/i.scitotenv.2003.12.005.
Lynch. MDJ; Bartram. AK; Neufeld. JD. (2012). Targeted recovery of novel phylogenetic
diversity from next-generation sequence data. ISME J 6: 2067-2077.
http://dx.doi.org/10.1038/ismei.2012.50.
Maaroufi, NI; Nordin, A; Palmqvist, K; Gundale, MJ. (2016). Chronic Nitrogen Deposition Has
a Minor Effect on the Quantity and Quality of Aboveground Litter in a Boreal Forest.
PLoS ONE 11: eO 162086. http://dx.doi.org/10.1371/iournal.pone.0162086.
Mabardv. RA: Waldbusser. GG: Conway. F: Olsen. CS. (2015). PERCEPTION AND
RESPONSE OF THE US WEST COAST SHELLFISH INDUSTRY TO OCEAN
ACIDIFICATION: THE VOICE OF THE CANARIES IN THE COAL MINE. J
Shellfish Res 34: 565-572. http://dx.doi.org/10.2983/035.034.0241.
MacAvoy. SW: Bulger. AJ. (1995). Survival of brook trout (Salvelinus fontinalis) embryos and
fry in streams of different acid sensitivity in Shenandoah National Park, USA. Water Air
Soil Pollut 85: 445-450. http://dx.doi.org/10.1007/BF0Q476869.
Macdougall. SE; Carrick. HJ; Dewalle. DR. (2008). Benthic algae in episodically acidified
Pennsylvania streams. Northeast Nat 15: 189-208. http://dx.doi.org/10.1656/lQ92-
6194(2008)15ri89:BAIEAP12.0.CO:2.
Mack. MC: Schuur. EA; Bret-Harte. MS: Shaver. GR; Chapin. FS. III. (2004). Ecosystem carbon
storage in arctic tundra reduced by long-term nutrient fertilization. Nature 431: 440-443.
http://dx.doi.org/10.1038/nature02887.
Mack. RN; Simberloff. D; Lonsdale. WM; Evans. H; Clout. M; Bazzaz. FA. (2000). Biotic
invasions: Causes, epidemiology, global consequences, and control. Ecol Appl 10: 689-
710. http://dx.doi.org/10.2307/2641039.
June 2018
R-137
DRAFT: Do Not Cite or Quote

-------
MACTEC (MACTEC Engineering and Consulting, Inc). (2010). Clean air status and trends
(CASTNET) - 2008 annual report [EPA Report], Washington, DC: U. S. Environmental
Protection Agency, Office of Air and Radiation.
http://nepis.epa. gov/Exe/ZyPURL.cgi?Dockev=P 100BZBZ.txt.
Maestre. FT: Ouero. JL; Gotelli. NJ: Escudero. A: Ochoa. V: Delgado-Baquerizo, M: Garcia-
Gomez. M: Bowker. MA: Soli veres. S: Escolar. C: Garcia-Palacios. P; Berdugo, M:
Valencia, E: Gozalo, B: Gallardo, A: Aguilera, L: Arredondo, T; Blones, J: Boeken, B:
Bran. D: Conceicao, AA: Cabrera. O: Chaieb, M: Derak, M: Eldridge, DJ: Espinosa, CI:
Florentino. A: Gaitan. J: Gatica. MG: Ghiloufi. W: Gomez-Gonzalez. S: Gutierrez. JR:
Hernandez. RM: Huang. X: Huber-Sannwald. E: Jankiu. M: Miriti. M: Monerris. J: Mau.
RL: Morici. E: Naseri. K: Ospina. A: Polo. V: Prina. A: Pucheta. E: Ramirez-Collantes.
DA: Romao. R: Tighe. M: Torres-Diaz. C: Val. J: Veiga. JP: Wang. D: Zaadv. E. (2012).
Plant species richness and ecosystem multifunctionality in global drylands. Science 335:
214-218. http://dx.doi.Org/10.l 126/science. 1215442.
Magenheimer. JF: Moore. TR: Chmura. GL: Daoust. RJ. (1996). Methane and carbon dioxide
flux from a macrotidal salt marsh, Bay of Fundy, New Brunswick. Estuaries 19: 139-145.
http://dx.doi.org/10.2307/1352658.
Magill. AH: Aber. JD: Currie. WS: Nadelhoffer. KJ: Martin. ME: McDowell. WH: Melillo. JM:
Steudler. P. (2004). Ecosystem response to 15 years of chronic nitrogen additions at the
Harvard Forest LTER, Massachusetts, USA. For Ecol Manage 196: 7-28.
http://dx.doi.Org/10.1016/i.foreco.2004.03.033.
Magnani. F; Mencuccini. M: Borghetti. M: Berbigier. P; Berninger. F; Delzon. S: Grelle. A:
Hari. P; Jarvis. PG: Kolari. P; Kowalski. AS: Lankreiier. H: Law. BE: Lindroth. A:
Loustau. D: Manca. G: Moncrieff. JB: Ravment. M; Tedeschi. V: Valentini. R: Grace. J.
(2007). The human footprint in the carbon cycle of temperate and boreal forests. Nature
447: 848-850. http://dx.doi.org/10.1038/nature05847.
Malama na Honu. (2015). Malamanahonu. Turtles in the Hawaiian islands. Available online at
http://malamanahonu.org/turtles.asp (accessed March 16, 2015).
Maletsika. PA: Nanos. GD: Stavroulakis. GG. (2015). Peach leaf responses to soil and cement
dust pollution. Environ Sci Pollut Res Int 22: 15952-15960.
http://dx.doi.org/10.1007/sll356-015-4821-z.
Malkonen. E: Derome. J: Fritze. H: Helmisaari. HS: Kukkola. M; Kyto. M; Saarsalmi. A:
Salemaa. M. (1999). Compensatory fertilization of Scots pine stands polluted by heavy
metals. Nutr Cycling Agroecosyst 55: 239-268.
http://dx.doi.Org/10.1023/A:1009851326584.
Mallev. DF: Chang. PSS. (1994). Assessing the health of a zooplankton community in a small
Precambrian Shield lake during recovery from experimental acidification. J Aquat Ecosys
Health 3: 273-286. http://dx.doi.org/10.1007/BF0Q115286.
Malm, WC: Rodriguez, MA: Schichtel, BA: Gebhart, KA: Thompson, TM: Barna, MG:
Benedict, KB: Carrico, CM: Collett, JL, Jr. (2016). A hybrid modeling approach for
estimating reactive nitrogen deposition in Rocky Mountain National Park. Atmos
Environ 126: 258-273. http://dx.doi.Org/10.1016/i.atmosenv.2015.l 1.060.
June 2018
R-138
DRAFT: Do Not Cite or Quote

-------
Malm. WC; Schichtel. BA; Barn a. MG; Gebhart KA; Rodriguez. MA; Collett JL. Jr; Carried.
CM; Benedict. KB; Prenni. A J; Kreidenweis. SM. (2013). Aerosol species concentrations
and source apportionment of ammonia at Rocky Mountain National Park. J Air Waste
Manag Assoc 63: 1245-1263. http://dx.doi.org/10.1080/10962247.2013.8Q4466.
Malone. TC; Conlev. DJ; Fisher. TR; Gilbert. PM; Harding. LW; Sellner. KG. (1996). Scales of
nutrient-limited phytoplankton productivity in Chesapeake Bay. Estuaries Coasts 19:
371-385. http://dx.doi.org/10.2307/1352457.
Mandvam. K; Jumpponen. A. (2008). Seasonal and temporal dynamics of arbuscular
mycorrhizal and dark septate endophytic fungi in a tallgrass prairie ecosystem are
minimally affected by nitrogen enrichment. Mycorrhiza 18: 145-155.
http://dx.doi.org/10.1007/s00572-008-Q165-6.
Mann. LK; Johnson. DW; West. DC; Cole. DW; Hornbeck. JW; Martin. CW; Riekerk. H; Smith.
CT; Swank. WT; Tritton. LM; Van Lear. DH. (1988). Effects on whole-tree and stem-
only clearcutting on postharvest hydrologic losses, nutrient capital, and regrowth. Forest
Sci 34: 412-428.
Manninen. S; Kivimaki. S; Leith. ID; Leeson. SR; Sheppard. LJ. (2016). Nitrogen deposition
does not enhance Sphagnum decomposition. Sci Total Environ 571: 314-322.
http://dx.doi.Org/10.1016/i.scitotenv.2016.07.152.
Manning. P; Saunders. M; Bardgett. RD; Bonkowski. M; Bradford. MA; Ellis. RJ; Kandeler. E;
Marhan. S; Tscherko. D. (2008). Direct and indirect effects of nitrogen deposition on
litter decomposition. Soil Biol Biochem 40: 688-698.
http://dx.doi.Org/10.1016/i.soilbio.2007.08.023.
Mant. RC; Jones. PL; Reynolds. B; Ormerod. SJ; Pullin. AS. (2013). A systematic review of the
effectiveness of liming to mitigate impacts of river acidification on fish and macro-
invertebrates [Review], Environ Pollut 179: 285-293.
http://dx.doi.Org/10.1016/i.envpol.2013.04.019.
Mao. J; Paulot. F; Jacob. DJ; Cohen. RC; Crounse. JD; Wennberg. PO; Keller. CA; Hudman.
RC; Barklev. MP; Horowitz. LW. (2013). Ozone and organic nitrates over the eastern
United States: Sensitivity to isoprene chemistry. J Geophys Res Atmos 118:1125611268.
http://dx.doi.org/10.1002/igrd.50817.
Marchenko. S; Krotkov. NA; Lamsal. LN; Celarier. EA; Swartz. WH; Bucsela. EJ. (2015).
Revising the slant column density retrieval of nitrogen dioxide observed by the ozone
monitoring instrument. J Geophys Res Atmos 120: 5670-5692.
http://dx.doi. org/10.1002/2014JD022913.
Margesin. R; Plaza. GA; Kasenbacher. S. (2011). Characterization of bacterial communities at
heavy-metal-contaminated sites. Chemosphere 82: 1583-1588.
http://dx.doi.Org/10.1016/i.chemosphere.2010.l 1.056.
Marklein. AR; Winbourne. J. ovB; Enders. SK; Gonzalez. DJX; van Huvsen. TL; Izquierdo. JE;
Light. PR; Liptzin. D; Miller. KE; Morford. SL; Norton. RA; Houlton. BZ. (2016).
Mineralization ratios of nitrogen and phosphorus from decomposing litter in temperate
versus tropical forests. Global Ecology and Biogeography 25: 335-346.
http://dx.doi.org/10. Ill 1/geb. 12414.
June 2018
R-139
DRAFT: Do Not Cite or Quote

-------
Marr, LC; Moore. TO; Klapmever, ME; Killar, MB. (2013). Comparison of NO(x) fluxes
measured by eddy covariance to emission inventories and land use. Environ Sci Technol
47: 1800-1808. http://dx.doi.org/10.1021/es303150y.
Marshall. KN; Kaplan. IC; Hodgson. EE; Hermann. A; Busch, DS; Mcelhany, P; Essington, TE;
Harvey. CJ; Fulton. EA. (2017). Risks of ocean acidification in the California Current
food web and fisheries: ecosystem model projections. Global Change Biol 23: 1525-
1539. http://dx.doi.org/10.Ill 1/gcb. 13594.
Martel. A; Blooi. M; Adriaensen. C; Van Rooii. P; Beukema. W; Fisher. MC; Farrer. RA;
Schmidt. BR; Tobler. U; Goka. K; Lips. KR; Muletz. C; Zamudio. KR; Bosch. J; Lotters.
S; Wombwell. E; Garner. TW; Cunningham. AA; Spitzen-Van der Sluiis. A; Salvidio. S;
Ducatelle, R; Nishikawa, K; Nguyen, TT; Kolby, JE; Van Bocxlaer, I; Bossuyt, F;
Pasmans. F. (2014). Wildlife disease. Recent introduction of a chytrid fungus endangers
Western Palearctic salamanders. Science 346: 630-631.
http://dx.doi.org/10.1126/science. 1258268.
Martin, JL; Hanke, AR; LeGresley, MM. (2009). Long term phytoplankton monitoring,
including harmful algal blooms, in the Bay of Fundy, eastern Canada. Journal of Sea
Research 61: 76-83. http://dx.doi.Org/10.1016/i.seares.2008.05.011.
Martin, KL; Goebel, PC. (2012). Decline in riparian Tsuga canadensis forests of the central
Appalachians across an Adelges tsugae invasion chronosequence. Torrey Botanical
Society Journal 139: 367-378. http://dx.doi.Org/10.3159/TORREY-D-12-00Q12.l.
Martin, MH; Coughtrev, PJ. (1981). Impact of metals on ecosystem function and productivity. In
NW Lepp (Ed.), Effect of heavy metal pollution on plants: Volume 2: Metals in the
environment (pp. 119-158). Barking, UK: Applied Science Publishers.
http://dx.doi.org/10.1007/978-94-009-8099-0 4.
Martin, RV; Chance, K; Jacob, DJ; Kurosu, TP; Spurr, RJD; Bucsela, E; Gleason, JF; Palmer,
PI; Bey, I; Fiore, AM; Li, O; Yantosca, RM; Koelemeiier, RBA. (2002). An improved
retrieval of tropospheric nitrogen dioxide from GOME. J Geophys Res 107: 9-1-9-21.
http://dx.doi.org/10.1029/2001JD001Q27.
Martin, RV; Sioris, CE; Chance, K; Ryerson, TB; Bertram, TH; Wooldridge, PJ; Cohen, RC;
Neuman, JA; Swanson, A; Flocke, FM. (2006). Evaluation of space-based constraints on
nitrogen oxide emissions with regional aircraft measurements over and downwind of
eastern North America. J Geophys Res 111: D15308.
http://dx.doi.org/10.1029/2005JD00668Q.
Marusenko, Y; Garcia-Pichel, F; Hall, SJ. (2015). Ammonia-oxidizing archaea respond
positively to inorganic nitrogen addition in desert soils. FEMS Microbiol Ecol 91: 1-11.
http://dx.doi.org/10.1093/femsec/fiu023.
Marusenko, Y; Huber, DP; Hall, SJ. (2013). Fungi mediate nitrous oxide production but not
ammonia oxidation in aridland soils of the southwestern US. Soil Biol Biochem 63: 24-
36. http://dx.doi.Org/10.1016/i.soilbio.2013.03.018.
June 2018
R-140
DRAFT: Do Not Cite or Quote

-------
Marvin-Dipasquale. M; Lutz. MA; Brigham. ME; Krabbenhoft. DP; Aiken. GR; Orem. WH;
Hall. BP. (2009). Mercury cycling in stream ecosystems. 2. Benthic methylmercury
production and bed sediment-pore water partitioning. Environ Sci Technol 43: 2726-
2732. http://dx.doi.org/10.1021/es802698v.
Marvin-Dipasquale. M; Windham-Myers. L; Agee. JL; Kakouros. E; Kieu. LH; Fleck. JA;
Alpers. CN; Strieker. CA. (2014). Methylmercury production in sediment from
agricultural and non-agricultural wetlands in the Yolo Bypass, California, USA. Sci Total
Environ 484: 288-299. http://dx.doi.Org/10.1016/i.scitotenv.2013.09.098.
Maskell. LC; Smart. SM; Bullock. JM; Thompson. K; Stevens. CJ. (2010). Nitrogen deposition
causes widespread loss of species richness in British habitats. Global Change Biol 16:
671-679. http://dx.doi.org/10.1111/j. 1365-2486.2009.02022.X.
Massad. RS; Nemitz. E; Sutton. MA. (2010). Review and parameterisation of bi-directional
ammonia exchange between vegetation and the atmosphere. Atmos Chem Phys 10:
10359-10386. http://dx.doi.org/10.5194/acp-10-10359-201Q.
Massicotte. R; Robidoux. PY; Sauve. S; Flipo. D; Fournier. M; Trottier. B. (2003). Immune
response of earthworms (Lumbricus terrestris, Eisenia andrei and Aporrectodea
tuberculata) following in situ soil exposure to atmospheric deposition from a cement
factory. J Environ Monit 5: 774-779. http://dx.doi.org/10.1039/b301956i.
Mast. MA; Campbell. DH; Ingersoll. GP; Foreman. WT; Krabbenhoft. DP. (2003). Atmospheric
deposition of nutrients, pesticides, and mercury in Rocky Mountain National Park,
Colorado, 2002. (Water Resources Investigations Report 03-4241). Reston, VA: U.S.
Department of the Interior, U.S. Geological Survey.
http://pubs.water.usgs.gov/wri034241.
Mast. MA; Clow. DW; Baron. JS; Wetherbee. GA. (2014). Links between N deposition and
nitrate export from a high-elevation watershed in the Colorado Front Range. Environ Sci
Technol 48: 14258-14265. http://dx.doi.org/10.1021/es502461k.
Mast. MA; Turk. JT; Clow. DW; Campbell. DH. (2011). Response of lake chemistry to changes
in atmospheric deposition and climate in three high-elevation wilderness areas of
Colorado. Biogeochemistry 103: 27-43. http://dx.doi.org/10.1007/slQ533-010-9443-4.
Matson. P; Lohse. KA; Hall. SJ. (2002). The globalization of nitrogen deposition: Consequences
for terrestrial ecosystems. Ambio 31: 113-119. http://dx.doi.org/10.1579/0Q44-7447-
31.2.113.
Matsui, T; Beltran-Przekurat, A; Niyogi, D, ev; Pielke, RA, Sr; Coughenour, M. (2008). Aerosol
light scattering effect on terrestrial plant productivity and energy fluxes over the eastern
United States. J Geophys Res 113: D14S14. http://dx.doi.org/10.1029/2007JD0Q9658.
Matthews. DA; Babcock. DB; Nolan. JG; Prestigiacomo. AR; Effler. SW; Driscoll. CT;
Todorova, SG; Kuhr, KM. (2013). Whole-lake nitrate addition for control of
methylmercury in mercury-contaminated Onondaga Lake, NY. Environ Res 125: 52-60.
http://dx.doi.Org/10.1016/i.envres.2013.03.011.
Matthews. JW; Peralta. AL; Soni. A; Baldwin. P; Kent. AD; Endress. AG. (2009). Local and
landscape correlates of non-native species invasion in restored wetlands. Ecography 32:
1031-1039. http://dx.doi.org/10.1111/i. 1600-0587.2009.05863.x.
June 2018
R-141
DRAFT: Do Not Cite or Quote

-------
Matuszek. JE; Beggs, GL. (1988). Fish species richness in relation to lake area, pH, and other
abiotic factors in Ontario lakes. Can J Fish Aquat Sci 45: 1931-1941.
http://dx.doi.Org/10.l 139/f88-225.
Maul din. RL. Ill: Berndt T; Sipila. M; Paasonen. P; Petai a. T; Kim. S: Kurt en. T; Stratmann. F:
Kenninen. VM: Kulmala. M. (2012). A new atmospherically relevant oxidant of sulphur
dioxide. Nature 488: 193-196. http://dx.doi.org/10.1038/naturel 1278.
May. JD; Burdette, SB: Gilliam. FS: MBeth, A. (2005). Interspecific divergence in foliar nutrient
dynamics and stem growth in a temperate forest in response to chronic nitrogen inputs.
Can J For Res 35: 1023-1030. http://dx.doi.org/10.1139/X05-036.
McBride. MB: Simon. T; Tam. G: Wharton. S. (2013). Lead and arsenic uptake by leafy
vegetables grown on contaminated soils: Effects of mineral and organic amendments.
Water Air Soil Pollut 224: 1378. http://dx.doi.org/10.1007/sll270-Q12-1378-z.
Mccabe, RM; Hickey, BM; Kudela, RM; Lefebvre, KA; Adams, NG; Bill, BP; Gulland, FM;
Thomson, RE: Cochlan, WP; Trainer, VL. (2016). An unprecedented coastwide toxic
algal bloom linked to anomalous ocean conditions. Geophys Res Lett 43: 10366-10376.
http ://dx. doi .org/10.1002/2016GL070023.
McCalley, CK; Sparks, JP. (2009). Abiotic gas formation drives nitrogen loss from a desert
ecosystem. Science 326: 837-840. http://dx.doi.Org/10.l 126/science. 1178984.
McCann, KS. (2000). The diversity-stability debate. Nature 405: 228-233.
http://dx.doi.org/10.1038/35012234.
McCarthy, MJ; Newell, SE; Carini, SA; Gardner, WS. (2015). Denitrification dominates
sediment nitrogen removal and is enhanced by bottom-water hypoxia in the Northern
Gulf of Mexico. Estuaries Coasts 38: 2279-2294. http://dx.doi.org/10.1007/sl2237-Q15-
9964-0.
McClain, ME: Bover, EW: Dent, CL; Gergel, SE: Grimm, NB; Groffman, PM; Hart, SC:
Harvey, JW: Johnston, CA; Mavorga, E; McDowell, WH: Pinav, G. (2003).
Biogeochemical hot spots and hot moments at the interface of terrestrial and aquatic
ecosystems. Ecosystems 6: 301-312. http://dx.doi.org/10.1007/slQ021-003-0161-9.
McColl, JG. (1981). Effects of acid rain on plants and soils in California. Berkeley, CA:
Califormia Air Resources Board.
McCormick, JH; Jensen, KM; Anderson, LE. (1989). Chronic effects of low pH and elevated
aluminum on survival, maturation, spawning, and embryo-larval development of
the fathead minnow in soft water. Water Air Soil Pollut 43: 293-307.
McCormick, SD; Keves, A: Nislow, KH; Monette, MY. (2009). Impacts of episodic acidification
on in-stream survival and physiological impairment of Atlantic salmon (Salmo salar)
smolts. Can J Fish Aquat Sci 66: 394-403. http://dx.doi.Org/10.l 139/F09-002.
McCrackin, ML: Elser, JJ. (2010). Atmospheric nitrogen deposition influences denitrification
and nitrous oxide production in lakes. Ecology 91: 528-539. http://dx.doi.org/10.1890/08-
2210.1.
June 2018
R-142
DRAFT: Do Not Cite or Quote

-------
McCrackin. ML; Elser. JJ. (2012). Denitrification kinetics and denitrifier abundances in
sediments of lakes receiving atmospheric nitrogen deposition (Colorado, USA).
Biogeochemistry 108: 39-54. http://dx.doi.ore/10.1007/sl0533-011-9571-5.
McCrackin. ML: Harrison. JA; Compton. JE. (2013). A comparison of NEWS and SPARROW
models to understand sources of nitrogen delivered to US coastal areas. Biogeochemistry
114: 281-297. http://dx.doi.org/10.1007/slQ533-012-9809-x.
McCrackin. ML: Harrison. JA: Compton. JE. (2014). Factors influencing export of dissolved
inorganic nitrogen by major rivers: A new, seasonal, spatially explicit, global model.
Global Biogeochem Cycles 28: 269-285. http://dx.doi.org/10.1002/2013GB0Q4723.
McCrackin. ML: Harrison. JA: Compton. JE. (2015). Future riverine nitrogen export to coastal
regions in the United States: Prospects for improving water quality. J Environ Qual 44:
345-355. http://dx.doi.org/10.2134/iea2014.02.0Q81.
McCune, B; Geiser, L. (1997). Macrolichens of the Pacific Northwest. In B McCune; L Geiser
(Eds.). Corvallis, OR: Oregon State University Press, USDA Forest Service.
Mccutchan. JH; Saunders. JF; Pribvl. AL; Lewis. WM. (2003). Open-channel estimation of
denitrification. Limnology and Oceanography: Methods 1: 74-81.
Mcdonald. DG. (1983). The effects of H+ upon the gills of freshwater fish. Can J Zool 61: 691-
703. http://dx.doi.Org/10.l 139/z83-093.
McDonnell, MJ; Pickett, STA; Groffman, P; Bohlen, P; Pouyat, RV; Zipperer, WC; Parmelee,
RW: Carreiro, MM: Medley, K. (1997). Ecosystem processes along an urban-to-rural
gradient. Urban Ecosyst 1: 21-36. http://dx.doi.Org/10.1023/A:1014359024275.
Mcdonnell, TC; Belvazid, S; Sullivan, TJ; Bell, M; Clark, C; Blett, T; Evans, T; Cass, W;
Hyduke, A; Sverdrup, H. (2018a). Vegetation dynamics associated with changes in
atmospheric nitrogen deposition and climate in hardwood forests of Shenandoah and
Great Smoky Mountains National Parks, USA. Environ Pollut 237: 662-674.
http://dx.doi.Org/10.1016/i.envpol.2018.01.l 12.
McDonnell, TC: Belvazid, S; Sullivan, TJ: Sverdrup, H; Bowman, WD: Porter, EM. (2014a).
Modeled subalpine plant community response to climate change and atmospheric
nitrogen deposition in Rocky Mountain National Park, USA. Environ Pollut 187: 55-64.
http://dx.doi.Org/10.1016/i.envpol.2013.12.021.
McDonnell, TC; Cosby, BJ; Sullivan, TJ. (2012). Regionalization of soil base cation weathering
for evaluating stream water acidification in the Appalachian Mountains, USA. Environ
Pollut 162: 338-344. http://dx.doi.Org/10.1016/i.envpol.2011.ll.025.
Mcdonnell, TC; Cosby, BJ; Sullivan, TJ; Mcnultv, SG; Cohen, EC. (2010). Comparison among
model estimates of critical loads of acidic deposition using different sources and scales of
input data. Environ Pollut 158: 2934-2939.
http://dx.doi.Org/10.1016/i.envpol.2010.06.007.
June 2018
R-143
DRAFT: Do Not Cite or Quote

-------
Mcdonnell, TC; Reinds, GJ; Sullivan. TJ; Clark. CM; Bonten, LTC; Mol-Dijkstra, JP;
Wamelink. GWW; Dovciak. M. (2018b). Feasibility of coupled empirical and dynamic
modeling to assess climate change and air pollution impacts on temperate forest
vegetation of the eastern United States. Environ Pollut 234: 902-914.
http://dx.doi.Org/10.1016/i.envpol.2017.12.002.
McDonnell. TC: Sloat. MR: Sullivan. TJ: Dolloff. CA; Hessburg, PF; Povak. NA; Jackson. WA;
Sams, C. (2015). Downstream warming and headwater acidity may diminish coldwater
habitat in southern Appalachian mountain streams. PLoS ONE 10: e0134757.
http://dx.doi.org/10.1371/iournal.pone.0134757.
McDonnell. TC: Sullivan. TJ: Cosby. BJ: Jackson. WA: Elliott. KJ. (2013). Effects of climate,
land management, and sulfur deposition on soil base cation supply in national forests of
the southern Appalachian Mountains. Water Air Soil Pollut 224: 1733.
http://dx.doi.org/10.1007/sll270-Q13-1733-8.
McDonnell. TC; Sullivan. TJ; Hessburg. PF; Reynolds. KM; Povak. NA; Cosby. BJ; Jackson.
W; Salter, RB. (2014b). Steady-state sulfur critical loads and exceedances for protection
of aquatic ecosystems in the U.S. Southern Appalachian Mountains. J Environ Manage
146: 407-419. http://dx.doi.Org/10.1016/i.ienvman.2014.07.019.
McDonough, AM; Watmough, SA. (2015). Impacts of nitrogen deposition on herbaceous ground
flora and epiphytic foliose lichen species in southern Ontario hardwood forests. Environ
Pollut 196: 78-88. http://dx.doi.Org/10.1016/i.envpol.2014.09.013.
McEathron, KM; Mitchell, MJ; Zhang, L. (2013). Acid-base characteristics of the Grass Pond
watershed in the Adirondack Mountains of New York State, USA: Interactions among
soil, vegetation and surface waters. Hydrol Earth Syst Sci 17: 2557-2568.
http://dx.doi.org/10.5194/hess-17-2557-2013.
McGill, BM; Sutton-Grier, AE; Wright, JP. (2010). Plant trait diversity buffers variability in
denitrification potential over changes in season and soil conditions. PLoS ONE 5:
el 1618. http://dx.doi.org/10.1371/iournal.pone.0011618.
Mchale, MR; Burns, DA; Siemion, J; Antidormi, MR. (2017). The response of soil and stream
chemistry to decreases in acid deposition in the Catskill Mountains, New York, USA.
Environ Pollut 229: 607-620. http://dx.doi.Org/10.1016/i.envpol.2017.06.001.
McHugh, TA; Morrissev, EM; Mueller, RC; Gallegos-Graves, LV; Kuske, CR; Reed, SC.
(2017). Bacterial, fungal, and plant communities exhibit no biomass or compositional
response to two years of simulated nitrogen deposition in a semiarid grassland. Environ
Microbiol 19: 1600-1611. http://dx.doi.Org/10.l 111/1462-2920.13678.
McLauchlan, KK; Craine, JM; Nippert, JB; Ocheltree, TW. (2014). Lack of eutrophication in a
tallgrass prairie ecosystem over 27 years. Ecology 95: 1225-1235.
http://dx.doi.Org/10.1890/13-1068.l.
McLaughlin, K; Sutula, M; Busse, L; Anderson, S; Crooks, J; Dagit, R; Gibson, D; Johnston, K;
Stratton, L. (2014). A regional survey of the extent and magnitude of eutrophi cation in
Mediterranean estuaries of Southern California, USA. Estuaries Coasts 37: 259-278.
http://dx.doi.org/10.1007/sl2237-013-967Q-8.
June 2018
R-144
DRAFT: Do Not Cite or Quote

-------
McLaughlin. SB; Tioelker. MJ. (1992). Growth and physiological changes in red spruce saplings
associated with acidic deposition at high elevations in the Southern Appalachians, USA.
For Ecol Manage 51: 43-51. http://dx.doi.org/10.1016/0378-1127(92)90470-T.
Mcleod. AR; Holland. MR; Shaw. PJA; Sutherland. PM; Darrall. NM; Skeffington. RA. (1990).
Enhancement of nitrogen deposition to forest trees exposed to S02. Nature 347: 277-279.
http://dx.doi.org/10.1038/347277a0.
McMurrav. JA; Roberts. DW; Fenn. ME; Geiser. LH; Jovan. S. (2013). Using epiphytic lichens
to monitor nitrogen deposition near natural gas drilling operations in the Wind River
Range, WY, USA. Water Air Soil Pollut 224. http://dx.doi.org/10.1007/s 11270-013-
1487-3.
McMurrav. JA; Roberts. DW; Geiser. LH. (2015). Epiphytic lichen indication of nitrogen
deposition and climate in the Northern Rocky Mountains, USA. Ecol Indicat 49: 154-
161. http://dx.doi.Org/10.1016/i.ecolind.2014.10.015.
McNicol. DK; Mallory. ML; Wedeles. CHR. (1995). Assessing biological recovery of acid-
sensitive lakes in Ontario, Canada. Water Air Soil Pollut 85: 457-462.
http://dx.doi.org/10.1007/BF0Q476871.
McNulty, SG; Aber, JD; Newman, SD. (1996). Nitrogen saturation in a high elevation New
England spruce-fir stand. For Ecol Manage 84: 109-121. http://dx.doi.org/10.1016/Q378-
1127(96^03742-5.
McNulty. SG; Boggs. J; Aber. JD; Rustad. L; Magill. A. (2005). Red spruce ecosystem level
changes following 14 years of chronic N fertilization. For Ecol Manage 219: 279-291.
http://dx.doi.Org/10.1016/i.foreco.2005.09.004.
McNulty. SG; Boggs. JL. (2010). A conceptual framework: Redefining forest soil's critical acid
loads under a changing climate. Environ Pollut 158: 2053-2058.
http://dx.doi.Org/10.1016/i.envpol.2009.l 1.028.
McNulty. SG; Cohen. EC; Myers. JAM; Sullivan. TJ; Li. H. (2007). Estimates of critical acid
loads and exceedances for forest soils across the conterminous United States. Environ
Pollut 149: 281-292. http://dx.doi.Org/10.1016/i.envpol.2007.05.025.
Mcquilling. AM; Adams. PJ. (2015). Semi-empirical process-based models for ammonia
emissions from beef, swine, and poultry operations in the United States. Atmos Environ
120: 127-136. http://dx.doi.Org/10.1016/i.atmosenv.2015.08.084.
McSkimming. C; Tanner. JE; Russell. BP; Connell. SD. (2015). Compensation of nutrient
pollution by herbivores in seagrass meadows. Exp Mar Bio Ecol 471: 112-118.
http://dx.doi.Org/10.1016/i.iembe.2015.05.018.
McWilliams, PG; Potts, WTW. (1978). The effects of pH and calcium concentrations on gill
potentials in the brown trout, Salmo trutta. J Comp Physiol [B] 126: 277-286.
http://dx.doi.org/10.1007/BF0Q688938.
Meador, MR; Carlisle, DM. (2007). Quantifying tolerance indicator values for common stream
fish species of the United States. Ecol Indicat 7: 329-338.
http://dx.doi.Org/10.1016/i.ecolind.2006.02.004.
June 2018
R-145
DRAFT: Do Not Cite or Quote

-------
Medeiros, JS; Tomeo, NJ; Hewins, CR; Rosenthal DM. (2016). Fast-growing Acer rubrum
differs from slow-growing Quercus alba in leaf, xylem and hydraulic trait coordination
responses to simulated acid rain. Tree Physiol 36: 1032-1044.
http://dx.doi.org/10.1093/treephvs/tpwQ45.
Meerhoff M; Teixeira-De Mello. F; Kruk. C: Alonso. C: Gonzalez-Bergonzoni, I; Pacheco. JP:
Lacerot G: Arim, M; Beklioglu, M: Brucet S: Govenola, G: Iglesias, C: Mazzeo, N:
Kosten, S; Jeppesen, E. (2012). Environmental warming in shallow lakes: A review of
potential changes in community structure as evidenced from space-for-time substitution
approaches. In U Jacob; G Woodward; EJ O'Gorman (Eds.), Adv Ecol Res (1st ed., pp.
259-349). Waltham, MA: Elsevier/Academic Press. http://dx.doi.org/10.1016/B978-Q-12-
396992-7.00004-6.
Meesenburg, H: Ahrends, B; Fleck. S: Wagner. M; Fortmann, H; Scheler, B; Klinck, U. we:
Dammann. I; Eichhorn. J: Mindrup. M; Meiwes. KJ. (2016). Long-term changes of
ecosystem services at Soiling, Germany: Recovery from acidification, but increasing
nitrogen saturation? Ecol Indicat 65: 103-112.
http://dx.doi.Org/10.1016/i.ecolind.2015.12.013.
Meier. S: Borie. F; Bolan. N: Corneio. P. (2012). Phytoremediation of Metal-Polluted Soils by
Arbuscular Mycorrhizal Fungi. Crit Rev Environ Sci Tech 42: 741-775.
http://dx.doi.org/10.1080/10643389.201Q.528518.
Meixner. T; Fenn. M. (2004). Biogeochemical budgets in a Mediterranean catchment with high
rates of atmospheric N depositionimportance of scale and temporal asynchrony.
Biogeochemistry 70: 331-356. http://dx.doi.org/10.1007/slQ533-003-4106-3.
Meixner, T; Fenn, ME; Wohlgemuth, P; Oxford, M; Riggan, P. (2006). N saturation symptoms
in chaparral catchments are not reversed by prescribed fire. Environ Sci Technol 40:
2887-2894. http://dx.doi.org/10.1021/es051268z.
Melack, JM; Cooper, SC: Jenkins, TM; Barmuta, L, Jr; Hamilton, S: Kratz, K; Sickman, J:
Soiseth, C. (1989). Chemical and biological characteristics of Emerald Lake and the
streams in its watershed, and the response of the lake and streams to acidic deposition.
Final report. (A6-184-32). Santa Barbara, CA: California Air Resources Board Resources
Agency, Univerisity of California, Santa Barbara.
Melack, JM; Stoddard, JL. (1991). Sierra Nevada: Unacidified, very dilute waters and mildly
acidic atmospheric deposition. In DF Charles; S Christie (Eds.), Acidic Deposition and
Aquatic Ecosystems (pp. 503-530). New York, NY: Springer-Verlag.
http://dx.doi.org/10.1007/978-l-4613-9038-l 21.
Melack, JM; Stoddard, JL; Ochs, CA. (1985). Major ion chemistry and sensitivity to acid
precipitation of Sierra Nevada lakes. Water Resour Res 21: 27-32.
http://dx.doi.org/10.1029/WR021i001pQ0027.
Meland, S: Borgstrem, R; Heier, LS: Rosseland, BO: Lindholm, O; Salbu, B. (2010). Chemical
and ecological effects of contaminated tunnel wash water runoff to a small Norwegian
stream. Sci Total Environ 408: 4107-4117.
http://dx.doi.Org/10.1016/i.scitotenv.2010.05.034.
June 2018
R-146
DRAFT: Do Not Cite or Quote

-------
Melillo. JM; Aber. J; Muratore. JF. (1982). Nitrogen and lignin control of hardwood leaf litter
decomposition dynamics. Ecology 63: 621-626. http://dx.doi.org/10.2307/193678Q.
Melliti Ben Garali. S: Sahraoui. I; de La Iglesia. P; Chalghaf M; Diogene. J: Ksouri. J: Sakka
Hlaili. A. (2016). Effects of nitrogen supply on Pseudo-nitzschia calliantha and Pseudo-
nitzschia cf. seriata: field and laboratory experiments. Ecotoxicology 25: 1211-1225.
http://dx.doi.org/10.1007/slQ646-016-1675-l.
Mem on. AR; Schroder. P. (2009). Implications of metal accumulation mechanisms to
phytoremediation. Environ Sci PollutRes Int 16: 162-175.
http://dx.doi.org/10.1007/sll356-008-0Q79-z.
Mendelssohn. IA. (1979). The influence of nitrogen level, form, and application method on the
growth response of Spartina alterniflora in North Carolina. Estuaries 2: 106-112.
http://dx.doi.org/10.2307/1351634.
Menon, M; Parratt, RT; Kropf, CA; Tyler, SW. (2010). Factors contributing to nitrate
accumulation in mesic desert vadose zones in Spanish Springs Valley, Nevada (USA). J
Arid Environ 74: 1033-1040. http://dx.doi.Org/10.1016/i.iaridenv.2010.03.005.
Mesquita. SR; van Drooge. BL; Barata. C: Vieira. N: Guimaraes. L; Pina. B. (2014a). Toxicity
of atmospheric particle-bound PAHs: an environmental perspective [Review], Environ
Sci PollutRes Int 21: 11623-11633. http://dx.doi.org/10.1007/s 11356-014-2628-v.
Mesquita. SR: van Drooge. BL: Oliveira. E; Grimalt. JO: Barata. C: Vieira. N: Guimaraes. L;
Pina. B. (2015). Differential embryotoxicity of the organic pollutants in rural andurban
air particles. Environ Pollut 206: 535-542.
http://dx.doi.Org/10.1016/i.envpol.2015.08.008.
Mesquita. SR: van Drooge. BL: Reche. C: Guimaraes. L; Grimalt. JO: Barata. C: Pina. B.
(2014b). Toxic assessment of urban atmospheric particle-bound PAHs: Relevance of
composition and particle size in Barcelona (Spain). Environ Pollut 184: 555-562.
http://dx.doi.Org/10.1016/i.envpol.2013.09.034.
Meunier. CL; Gundale. MJ; Sanchez. IS: Liess. A. (2016). Impact of nitrogen deposition on
forest and lake food webs in nitrogen-limited environments [Review], Global Change
Biol 22: 164-179. http://dx.doi.org/10.1111/gcb. 12967.
Mever-Gruenefeldt. M: Belz. K: Calvo. L: Marcos. E: von Oheimb. G: Haerdtle. W. (2016).
Marginal Calluna populations are more resistant to climate change, but not under high-
nitrogen loads. Plant Ecol 217: 111-122. http://dx.doi.org/10.1007/sl 1258-015-0563-8.
Meyer. C: Bernard. N: Moskura. M: Toussaint. ML: Denaver. F; Gilbert. D. (2010a). Effects of
urban particulate deposition on microbial communities living in bryophytes: An
experimental study. Ecotoxicol Environ Saf 73: 1776-1784.
http://dx.doi.Org/10.1016/i.ecoenv.2010.07.012.
Meyer. C: D.gilbert: Gillet. F; Moskura. M; Franchi. M: Bernard. N. (2012). Using bryophytes
and their associated testate amoeba microsystems as indicators of atmospheric pollution.
Ecol Indicat 13: 144-151. http://dx.doi.Org/10.1016/i.ecolind.2011.05.020.
June 2018
R-147
DRAFT: Do Not Cite or Quote

-------
Meyer. C; Desalme, D; Bernard. N; Binet P; Toussaint ML; Gilbert. D. (2013). Using testate
amoeba as potential biointegrators of atmospheric deposition of phenanthrene (polycyclic
aromatic hydrocarbon) on "moss/soil interface-testate amoeba community"
microecosystems. Ecotoxicology 22: 287-294. http://dx.doi.org/10.1007/slQ646-012-
1025-x.
Meyer. C: Gilbert. D; Gaudry. A: Franchi. M; Nguven-Viet. H; Fabure. J: Bernard. N. (2010b).
Relationship of atmospheric pollution characterized by gas (N02) and particles (PM10)
to microbial communities living in bryophytes at three differently polluted sites (rural,
urban, and industrial). Microb Ecol 59: 324-334. http://dx.doi.org/10.1007/sQ0248-009-
9580-2.
Michelena, TM; Farrell, JL; Winkler, DA; Goodrich, CA; Boylen, CW; Sutherland, JW;
Nierzwicki-Bauer, SA. (2016). Aluminum toxicity risk reduction as a result of reduced
acid deposition in Adirondack lakes and ponds. Environ Monit Assess 188: 636.
http://dx.doi.org/10.1007/slQ661-016-5589-4.
Middelburg, JJ; Levin, LA. (2009). Coastal hypoxia and sediment biogeochemistry.
Biogeosciences 6: 1273-1293. http://dx.doi.org/10.5194/bg-6-1273-2009.
Midglev, MG; Phillips, RP. (2016). Resource stoichiometry and the biogeochemical
consequences of nitrogen deposition in a mixed deciduous forest. Ecology 97: 3369-
3378. http://dx.doi.org/10.1002/ecy.1595.
Mikaelvan, AS; Zatsepin, AG; Chasovnikov, VK. (2013). Long-term changes in nutrient supply
of phytoplankton growth in the Black Sea. J Mar Syst 117-118: 53-64.
http://dx.doi.Org/10.1016/i.imarsvs.2013.02.012.
Mikkelson, KM; Bearup, LA; Maxwell, RM; Stednick, JD; Mccrav, JE; Sharp, JO. (2013). Bark
beetle infestation impacts on nutrient cycling, water quality and interdependent
hydrological effects. Biogeochemistry 115: 1-21. http://dx.doi.org/10.1007/slQ533-013-
9875-8.
Mikryukov, VS; Dulva, OV; Vorobeichik, EL. (2015). Diversity and spatial structure of soil
fungi and arbuscular mycorrhizal fungi in forest litter contaminated with copper smelter
emissions. Water Air Soil Pollut 226: 114. http://dx.doi.org/10.1007/sl 1270-014-2244-v.
Miller, DE; Watmough, SA. (2009). Air pollution, climate, soil acidity, and indicators of forest
health in Ontario's sugar maple forests. Can J For Res 39: 2065-2079.
http://dx.doi.Org/10.l 139/x09-124.
Miller, DJ; Sun, K; Tao, L, i; Pan, D; Zondlo, MA; Nowak, JB; Liu, Z; Diskin, G; Sachse, G;
Beversdorf, A; Ferrare, R; Scarino, AJ. (2015). Ammonia and methane dairy emission
plumes in the San Joaquin Valley of California from individual feedlot to regional scales.
J Geophys Res Atmos 120: 9718-9738. http://dx.doi.org/10.1002/2015JDQ23241.
Miller, MA; Kudela, RM; Mekebri, A; Crane, D; Oates, SC; Tinker, MT; Staedler, M; Miller,
WA; Toy-Choutka, S; Dominik, C; Hardin, D; Langlois, G; Murray, M; Ward, K; Jessup,
DA. (2010). Evidence for a novel marine harmful algal bloom: Cyanotoxin (microcystin)
transfer from land to sea otters. PLoS ONE 5: el2576.
http://dx.doi.org/10.1371/iournal.pone.0012576.
June 2018
R-148
DRAFT: Do Not Cite or Quote

-------
Miller. MP; Mcknight DM. (2015). Limnology of the Green Lakes Valley: phytoplankton
ecology and dissolved organic matter biogeochemistry at a long-term ecological research
site. Plant Ecol Divers 8: 689-702. http://dx.doi.org/10.1080/17550874.2Q12.738255.
Millett J: Foot GW: Svensson, BM. (2015). Nitrogen deposition and prey nitrogen uptake
control the nutrition of the carnivorous plant Drosera rotundifolia. Sci Total Environ 512:
631-636. http://dx.doi.Org/10.1016/i.scitotenv.2015.01.067.
Millett J: Svensson. BM: Newton. J: Rydin. H. (2012). Reliance on prey-derived nitrogen by the
carnivorous plant Drosera rotundifolia decreases with increasing nitrogen deposition.
NewPhytol 195: 182-188. http://dx.doi.Org/10.llll/i.1469-8137.2012.04139.x.
Milner. AM: Woodward. A: Freilich. JE; Black. RW: Resh. VH. (2016). Detecting significant
change in stream benthic macroinvertebrate communities in wilderness areas. Ecol
Indicat 60: 524-537. http://dx.doi.Org/10.1016/i.ecolind.2015.07.025.
Min, K, -E; Pusede, SE; Browne, EC: LaFranchi, BW: Wooldridge, PJ; Cohen, RC. (2014).
Eddy covariance fluxes and vertical concentration gradient measurements of NO and
N02 over a ponderosa pine ecosystem: Observational evidence for within-canopy
chemical removal of Nox. Atmos Chem Phys 14: 5495-5512.
http://dx.doi.org/10.5194/acp-14-5495-2014.
Mineau, MM: Fatemi, FR; Fernandez, IJ; Simon, KS. (2014). Microbial enzyme activity at the
watershed scale: Response to chronic nitrogen deposition and acute phosphorus
enrichment. Biogeochemistry 117: 131-142. http://dx.doi.org/10.1007/slQ533-013-9869-
6.
Minnich, RA; Bahre, CJ. (1995). Wildland fire and chaparral succession along the California
Baja-California boundary. International Journal of Wildland Fire 5: 13-24.
http://dx.doi.org/10.1071/WF9950013.
Minnich, RA; Barbour, MG: Burk, JH; Fernau, RF. (1995). Sixty years of change in California
conifer forests of the San Bernardino mountains. Conserv Biol 9: 902-914.
http://dx.doi.Org/10.1046/i.1523-1739.1995.09040902.x.
Minocha, R; Long, S: Thangavel, P; Minocha, SC: Eagar, C: Driscoll, CT. (2010). Elevation
dependent sensitivity of northern hardwoods to Ca addition at Hubbard Brook
Experimental Forest, NH, USA. For Ecol Manage 260: 2115-2124.
http://dx.doi.Org/10.1016/i.foreco.2010.09.002.
Minocha, R; Shortle, WC: Lawrence, GB; David, MB: Minocha, SC. (1997). Relationships
among foliar chemistry, foliar polyamines, and soil chemistry in red spruce trees growing
across the northeastern United States. Plant Soil 191: 109-122.
http://dx.doi.Org/10.1023/A:1004293523185.
Mitchell, CP J: Branfireun, BA; Kolka, RK. (2008a). Assessing sulfate and carbon controls on net
methylmercury production in peatlands: An in situ mesocosm approach. Appl Geochem
23: 503-518. http://dx.doi.Org/10.1016/i.apgeochem.2007.12.020.
Mitchell, CP J: Branfireun, BA: Kolka, RK. (2008b). Spatial characteristics of net methylmercury
production hot spots in peatlands. Environ Sci Technol 42: 1010-1016.
http://dx.doi.org/10.1021/es0704986.
June 2018
R-149
DRAFT: Do Not Cite or Quote

-------
Mitchell. CP J; Branfireun. BA; Kolka. RK. (2008c). Total mercury and methylmercury
dynamics in upland-peatland watersheds during snowmelt. Biogeochemistry 90: 225-241.
http://dx.doi.org/10.1007/slQ533-008-9246-z.
Mitchell CP J: Branfireun. BA: Kolka. RK. (2009). Methylmercury dynamics at the upland-
peatland interface: Topographic and hydrogeochemical controls. Water Resour Res 45.
http://dx.doi.org/10.1029/2008WR0Q6832.
Mitchell. CPJ: Gilmour. CC. (2008). Methylmercury production in a Chesapeake Bay salt marsh.
J Geophys Res 113. http://dx.doi.org/10.1029/2008JG00Q765.
Mitchell, MJ. (2011). Nitrate dynamics of forested watersheds: Spatial and temporal patterns in
North America, Europe and Japan. J For Res 16: 333-340.
http ://dx. doi .org/10.1007/s 10310-011 -0278-1.
Mitchell. MJ: Driscoll. CT; Kahl. JS: Likens. GE; Murdoch. PS: Pardo. LH. (1996). Climatic
control of nitrate loss from forested watersheds in the northeast United States. Environ
Sci Technol 30: 2609-2612. http://dx.doi.org/10.1021/es9600237.
Mitchell. MJ: Driscoll. CT: McHale. PJ: Roy. KM: Dong. Z. (2013). Lake/watershed sulfur
budgets and their response to decreases in atmospheric sulfur deposition: Watershed and
climate controls. Hydrolog Process 27: 710-720. http://dx.doi.org/10.1002/hyp.9670.
Mitchell. MJ: Likens. GE. (2011). Watershed sulfur biogeochemistry: Shift from atmospheric
deposition dominance to climatic regulation. Environ Sci Technol 45: 5267-5271.
http://dx.doi.org/10.1021/es200844n.
Mitchell. MJ: Lovett. G: Bailey. S: Beall. F; Burns. D: Buso. D: Clair. TA: Courchesne. F;
Duchesne. L: Eimers. C: Fernandez. I; Houle. D: Jeffries. DS: Likens. GE: Moran. MP:
Rogers. C: Schwede. D: Shanlev. J: Weathers. KC: Vet. R. (2011). Comparisons of
watershed sulfur budgets in southeast Canada and northeast US: New approaches and
implications. Biogeochemistry 103: 181-207. http://dx.doi.org/10.1007/slQ533-010-
9455-0.
Mitchell. RJ: Urpeth. HM; Britton. AJ: Black. H: Taylor. AR. (2016). Relative importance of
local- and large-scale drivers of alpine soil microarthropod communities. Oecologia 182:
913-924. http://dx.doi.org/10.1007/sQ0442-016-3688-0.
Mitsch. WJ: Gosselink. JG. (1986). Some other books of interest: Wetlands. Science 233: 1209.
http://dx.doi.Org/10.l 126/science.233.4769.1209-a.
Mitsch. WJ: Gosselink. JG. (2000). Wetlands (3rd ed.). New York, NY: Wiley & Sons, Inc.
Moe. SJ: Schartau. AK; Baekken. T; Mcfarland. B. (2010). Assessing macroinvertebrate metrics
for classifying acidified rivers across northern Europe. Freshw Biol 55: 1382-1404.
http://dx.doi.org/10.111 l/i.l365-2427.2010.02413.x.
Moin. NS: Nelson. KA: Bush. A: Bernhard. AE. (2009). Distribution and diversity of archaeal
and bacterial ammonia oxidizers in salt marsh sediments. Appl Environ Microbiol 75:
7461-7468. http://dx.doi.org/10.1128/AEM.01001-Q9.
Moller. D. (2014). Chemistry of the climate system. In Chemistry of the Climate System (2nd
ed.). Berlin, Germany: Walter de Gruyter GmbH.
June 2018
R-150
DRAFT: Do Not Cite or Quote

-------
Mom en. B; Lawrence. GB; Nierzwicki-Bauer. SA; Sutherland. JW; Eichler. LW; Harrison. JP;
Boylen. CW. (2006). Trends in summer chemistry linked to productivity in lakes
recovering from acid deposition in the Adirondack region of New York. Ecosystems 9:
1306-1317. http://dx.doi.org/10.1007/sl0021-006-0Q12-6.
Monette. MY; Biornsson. BT; McCormick. SD. (2008). Effects of short-term acid and aluminum
exposure on the parr-smolt transformation in Atlantic salmon (Salmo salar): Disruption
of seawater tolerance and endocrine status. Gen Comp Endocrinol 158: 122-130.
http://dx.doi.Org/10.1016/i.ygcen.2008.05.014.
Monette. MY: McCormick. SD. (2008). Impacts of short-term acid and aluminum exposure on
Atlantic salmon (Salmo salar) physiology: A direct comparison of parr and smolts. Aquat
Toxicol 86: 216-226. http://dx.doi.Org/10.1016/i.aquatox.2007.l 1.002.
Monette. MY: Yada. T; Matey. V: McCormick. SD. (2010). Physiological, molecular, and
cellular mechanisms of impaired seawater tolerance following exposure of Atlantic
salmon, Salmo salar, smolts to acid and aluminum. Aquat Toxicol 99: 17-32.
http://dx.doi.Org/10.1016/i.aquatox.2010.03.016.
Monteith. DT; Stoddard. JL; Evans. CD: de Wit. HA: Forsius. M; Etogasen. T; Wilander. A:
Skielkvale. BL; Jeffries. DS: Vuorenmaa. J: Keller. B; Kopacek. J: Veselv. J. (2007).
Dissolved organic carbon trends resulting from changes in atmospheric deposition
chemistry. Nature 450: 537-540. http://dx.doi.org/10.1038/nature06316.
Montross. GG: McGlynn. BL: Montross. SN: Gardner. KK. (2013). Nitrogen production from
geochemical weathering of rocks in southwest Montana, USA. 118: 1068-1078.
http://dx.doi.org/10.1002/igrg.20Q85.
Moore. J. -D; Houle. D. (2013). Soil and sugar maple response to 8 years of NH4N03 additions
in a base-poor northern hardwood forest. For Ecol Manage 310: 167-172.
http://dx.doi.Org/10.1016/i.foreco.2013.08.020.
Moore. J. -D; Ouimet. R; Duchesne. L. (2012). Soil and sugar maple response 15 years after
dolomitic lime application. For Ecol Manage 281: 130-139.
http://dx.doi.Org/10.1016/i.foreco.2012.06.026.
Moore, RB; Johnston, CM; Smith, RA; Mil stead, B. (2011). Source and delivery of nutrients to
receiving waters in the northeastern and mid-atlantic regions of the United States. J Am
Water Resour Assoc 47: 965-990. http://dx.doi.org/10.1111/i. 1752-1688.2011.00582.x.
Moore, T; Blodau, C: Turunen, J: Roulet, N: Richard, PJH. (2004). Patterns of nitrogen and
sulfur accumulation and retention in ombrotrophic bogs, eastern Canada. Global Change
Biol 11: 356-367. http://dx.doi.org/10.1111/i. 1365-2486.2004.00882.X.
Moran, R; Harvey, I: Moss, B; Feuchtmayr, H; Hatton, K; Heves, T; Atkinson, D. (2010).
Influence of simulated climate change and eutrophication on three-spined stickleback
populations: A large scale mesocosm experiment. Freshw Biol 55: 315-325.
http://dx.doi.org/10.1111/i. 1365-2427.2009.02276.X.
Morford, SL; Houlton, BZ; Dahlgren, RA. (2011). Increased forest ecosystem carbon and
nitrogen storage from nitrogen rich bedrock. Nature 477: 78-81.
http://dx.doi.org/10.1038/naturel0415.
June 2018
R-151
DRAFT: Do Not Cite or Quote

-------
Morier, I; Guenat C; Sieewolf, R; Vedv, J. -C; Schleppi, P. (2008). Dynamics of atmospheric
nitrogen deposition in a temperate calcareous forest soil. J Environ Qual 37: 2012-2021.
http://dx.doi.ore/10.2134/ieq2007.0563.
Morris. JL; le Roux. PC: Macharia. AN: Brunei le. A: Hebertson. EG: Lundeen. ZJ. (2013).
Organic, elemental, and geochemical contributions to lake sediment deposits during
severe spruce beetle (Dendroctonus rufipennis) disturbances. For Ecol Manage 289: 78-
89. http://dx.doi.ore/10.1016/i.foreco.2012.10.004.
Morris. JT. (1991). Effects of nitrogen loading on wetland ecosystems with particular reference
to atmospheric deposition. Annu Rev Ecol Evol Systemat 22: 257-279.
http://dx.doi.ore/10.1146/annurev.es. 22.110191.001353.
Morris. K. (2016). 2014 data summary of wet nitrogen deposition at Rocky Mountain National
Park. (Natural Resource Data Series NPS/NRSS/ARD/NRDS2016/1036). Denver, CO:
U.S. National Park Service.
Morris. K; Mast. A: Clow. D; Wetherbee. G: Baron. J: Taipale. C: Blett. T; Gay. D; Heath. J.
(2014). 2012 Monitoring and tracking wet nitrogen deposition at Rocky Mountain
National Park. (Natural Resource Report NPS/NRSS/ARD/NRR2014/757). Denver, CO:
U.S. National Park Service, https://www.nature.nps.eov/air/pubs/pdf/rmnp-trends/rmnp-
trends 2012.pdf.
Morrison. EW: Frev. SD; Sadowsky. JJ; van Diepen. LTA; Thomas. WK; Prinele. A. (2016).
Chronic nitrogen additions fundamentally restructure the soil fungal community in a
temperate forest. Fungal Ecology 23: 48-57.
http://dx.doi.ore/10.1016/i.funeco.2016.05.011.
Morrison. G: Greenine. HS: Yates. KK. (2011). Management Case Study: Tampa Bay, Florida.
31-76.
Morrissey, EM; Jenkins, AS: Brown, BL; Franklin, RB. (2013). Resource availability effects on
nitrate-reducing microbial communities in a freshwater wetland. Wetlands 33: 301-310.
http://dx.doi.ore/10.1007/sl3157-013-0384-2.
Morse, JL: Duran, J: Beall, F; Enanea, EM: Creed, IF: Fernandez, I; Groffman, PM. (2015a).
Soil denitrification fluxes from three northeastern North American forests across a range
of nitrogen deposition. Oecologia 177: 17-27. http://dx.doi.ore/10.1007/s00442-014-
3117-1.
Morse, JL: Duran, J: Groffman, PM. (2015b). Soil denitrification fluxes in a northern hardwood
forest: The importance of snowmelt and implications for ecosystem N budgets.
Ecosystems 18: 520-532. http://dx.doi.ore/10.1007/sl0021-015-9844-2.
Moseman-Valtierra, S: Gonzalez, R: Kroeeer, KD: Tane, J: Chao, W: Crusius, J: Bratton, J:
Green, A: Shelton, J. (2011). Short-term nitrogen additions can shift a coastal wetland
from a sink to a source of N20. Atmos Environ 45: 4390-4397.
http://dx.doi.ore/10.1016/i.atmosenv.2011.05.046.
Mosier, AC: Francis, CA. (2010). Denitrifier abundance and activity across the San Francisco
Bay estuary. Environ Microbiol Rep 2: 667-676. http://dx.doi.ore/10.111 l/i.1758-
2229.2010.00156.x.
June 2018
R-152
DRAFT: Do Not Cite or Quote

-------
Moslev. LM. (2015). Drought impacts on the water quality of freshwater systems; Review and
integration. Earth Sci Rev 140: 203-214.
http://dx.doi.ore/10.1016/j.earscirev. 2014.11.010.
Mosscrop. LE; Paterson. AM: DeSellas. AM: Kurek. J: Weeber. R: Smol, JP. (2015). Long-term
stability of cladoceran assemblages in small, shallow, Canadian Shield lakes experiencing
marked calcium declines. Aquat Sci 77: 547-561. http://dx.doi.org/10.1007/sQ0027-015-
0402-y.
Most of a. KMG: Liu. CO: Zhai. W: Minella. M; Vione. D: Gao. K: Minakata. D: Arakaki. T;
Yoshioka. T; Havakawa. K: Konohira. E: Tanoue. E: Akhand. A: Chan da. A: Wang. B;
Sakugawa. H. (2016). Reviews and Syntheses: Ocean acidification and its potential
impacts on marine ecosystems. Biogeosciences 13: 1767-1786.
http://dx.doi.org/10.5194/bg-13-1767-2016.
Mozdzer. TJ; Langlev. JA; Mueller. P; Megonigal. JP. (2016). Deep rooting and global change
facilitate spread of invasive grass. Biological Invasions 18: 2619-2631.
http://dx.doi.org/10.1007/slQ530-016-1156-8.
MPCA (Minnesota Pollution Control Agency). (2015a). Protecting wild rice from excess sulfate.
Saint Paul, MN. https://www.pca.state.mn.us/sites/default/files/wq-s6-43k.pdf.
MPCA (Minnesota Pollution Control Agency). (2015b). Protecting wild rice from too much
sulfate: DRAFT proposal. Available online at (accessed
Mueller. RC: Gallegos-Graves. L; Zak. PR: Kuske. CR. (2016). Assembly of Active Bacterial
and Fungal Communities Along a Natural Environmental Gradient. Microb Ecol 71: 57-
67. http://dx.doi.org/10.1007/sQ0248-015-0655-v.
Mukheriee. M; Ray. A: Post. AF; Mckav. RM; Bulleriahn. GS. (2016). Identification,
enumeration and diversity of nitrifying planktonic archaea and bacteria in trophic end
members of the Laurentian Great Lakes. J Great Lakes Res 42: 39-49.
http://dx.doi.Org/10.1016/i.iglr.2015.l 1.007.
Mulholland. PJ: Hall. RO. Jr; Sobota. DJ: Dodds. WK; Findlav. SEG: Grimm. NB; Hamilton.
SK; McDowell. WH; O'Brien. JM; Tank. JL; Ashkenas. LR; Cooper. L; Dahm. CN:
Gregory, SV; Johnson, SL; Meyer, JL; Peterson, BJ; Poole, GC; Valett, HM; Webster, J.
R.; Arango, CP: Beaulieu, JJ; Bernot, MJ; Burgin, A: Crenshaw, CL; Helton, AM:
Johnson, LT; Niederlehner, BR: Potter, JD; Sheiblev, RW: Thomas, SM. (2009). Nitrate
removal in stream ecosystems measured by 15N addition experiments: Denitrification.
Limnol Oceanogr 54: 666-680. http://dx.doi.Org/10.4319/lo.2009.54.3.0666.
Mulholland, PJ: Helton, AM: Poole, GC: Hall, RO: Hamilton, SK: Peterson, BJ: Tank, JL:
Ashkenas, LR: Cooper, LW: Dahm, CN: Dodds, WK: Findlav, SEG: Gregory, SV:
Grimm, NB: Johnson, SL: Mcdowell, WH: Meyer, JL: Valett, HM: Webster, JR: Arango,
CP: Beaulieu, JJ: Bernot, MJ: Burgin, AJ: Crenshaw, CL: Johnson, LT: Niederlehner,
BR: O'Brien, JM: Potter, JD: Sheiblev, RW: Sobota, DJ: Thomas, SM. (2008). Stream
denitrification across biomes and its response to anthropogenic nitrate loading. Nature
452: 202-206. http://dx.doi.org/10.1038/nature06686.
June 2018
R-153
DRAFT: Do Not Cite or Quote

-------
Mullen B; Touraine. B; Rennenberg, H. (1996). Interaction between atmospheric and
pedospheric nitrogen nutrition in spruce (Picea abies L. Karst) seedlings. Plant Cell
Environ 19: 345-355. http://dx.doi.Org/10.l 111/j. 1365-3040.1996.tb00257.x.
Mur. LA: Simpson. C: Kumari. A: Gupta. AK: Gupta. KJ. (2016). Moving nitrogen to the centre
of plant defence against pathogens [Review], Ann Bot 119: 703-709.
http://dx.doi.org/10.1093/aob/mcwl79.
Murdoch. PS: Burns. DA: Lawrence. GB. (1998). Relation of climate change to the acidification
of surface waters by nitrogen deposition. Environ Sci Technol 32: 1642-1647.
http://dx.doi.org/10.1021/es9708631.
Murdoch. PS: Stoddard. JL. (1992). The role of nitrate in the acidification of streams in the
Catskill Mountains of New York. Water Resour Res 28: 2707-2720.
http://dx.doi.org/10.1029/92WR00953.
Murphy, CA; Rose, KA; Rahman, MS: Thomas, P. (2009). Testing and applying a fish
vitellogenesis model to evaluate laboratory and field biomarkers of endocrine disruption
in Atlantic croaker (Micropogonias undulatus) exposed to hypoxia. Environ Toxicol
Chem 28: 1288-1303. http://dx.doi.Org/10.1897/08-304.l.
Murphy, CA; Thompson, PL; Vinebrooke, RD. (2010). Assessing the sensitivity of alpine lakes
and ponds to nitrogen deposition in the Canadian Rocky Mountains. Hydrobiologia 648:
83-90. http://dx.doi.org/10.1007/sl0750-010-Q146-6.
Murphy, GEP; Romanuk, TN. (2016). Data gaps in anthropogenically driven local-scale species
richness change studies across the Earth's terrestrial biomes. Ecol Evol 6: 2938-2947.
http://dx.doi.org/10.1002/ece3.20Q4.
Murphy, JF; Davv-Bowker, J: McFarland, B; Ormerod, SJ. (2013). A diagnostic biotic index for
assessing acidity in sensitive streams in Britain. Ecol Indicat 24: 562-572.
http://dx.doi.Org/10.1016/i.ecolind.2012.08.014.
Murphy, JF: Winterbottom, JH; Orton, S: Simpson, GL; Shilland, EM: Hildrew, AG. (2014).
Evidence of recovery from acidification in the macroinvertebrate assemblages of UK
fresh waters: A 20-year time series. Ecol Indicat 37: 330-340.
http://dx.doi.Org/10.1016/i.ecolind.2012.07.009.
Murphy, RR: Kemp, WM: Ball, WP. (2011). Long-term trends in Chesapeake Bay seasonal
hypoxia, stratification, and nutrient loading. Estuaries Coasts 34: 1293-1309.
http://dx.doi.org/10.1007/sl2237-011-9413-7.
Myking, T, or: Aarrestad, P; Derome, J: Bakkestuen, V: Bierke, JW: Gvtarsky, M: Isaeva, L:
Karaban, R: Korotkov, V: Lindgren, M: Lindroos, AJ: Rosberg, I; Salemaa, M:
Tommervik, H: Vassilieva, N. (2009). Effects of air pollution from a nickel-copper
industrial complex on boreal forest vegetation in the joint Russian-Norwegian-Finnish
border area. Boreal Environmental Research 14: 279-296.
Myles, L: Heuer, MW: Meyers, TP: Hovett, ZJ. (2012). A comparison of observed and
parameterized S02 dry deposition over a grassy clearing in Duke Forest. Atmos Environ
49: 212-218. http://dx.doi.Org/10.1016/i.atmosenv.2011.11.059.
June 2018
R-154
DRAFT: Do Not Cite or Quote

-------
Nadelhoffer. K; Downs. M; Fry. B; Magill. A; Aber. J. (1999a). Controls on N retention and
exports in a forested watershed. Environ Monit Assess 55: 187-210.
http://dx.doi.Org/10.1023/A:1006190222768.
Nadelh offer. KJ. (2000). The potential effects of nitrogen deposition on fine-root production in
forest ecosystems. NewPhytol 147: 131-139. http://dx.doi.org/10.1046/i. 1469-
8137.2000.00677.x.
Nadelh offer. KJ: Emmett BA; Gundersen. P; Kionaas. OJ; Koopmans. CJ: Schleppi. P; Tietema.
A: Wright RF. (1999b). Nitrogen deposition makes a minor contribution to carbon
sequestration in temperate forests. Nature 398: 145-148. http://dx.doi.org/10.1038/182Q5.
NADP (National Atmospheric Deposition Program). (2016). Total deposition 2015 Annual map
summary. NADP data report 2016-02 [EPA Report], Urbana-Champaign, IL: University
of Illinois at Urbana-Champaign.
Naeem, S; Li, SB. (1997). Biodiversity enhances ecosystem reliability. Nature 390: 507-509.
Naess. L. (2016). Summary of monitoring data: September 2016. Available online at (accessed
Naidoo. G; Chirkoot, D. (2004). The effects of coal dust on photosynthetic performance of the
mangrove, Avicennia marina in Richards Bay, South Africa. Environ Pollut 127: 359-
366.
Nair. RK; Perks. MP: Mencuccini. M. (2017). Decomposition nitrogen is better retained than
simulated deposition from mineral amendments in a temperate forest. Global Change
Biol 23: 1711-1724. http://dx.doi.Org/10.l 111/gcb. 13450.
Nair. RKF; Perks. MP: Weatherall. A: Baggs. EM: Mencuccini. M. (2016). Does canopy
nitrogen uptake enhance carbon sequestration by trees? Global Change Biol 22: 875-888.
http://dx.doi.org/10. Ill 1/gcb. 13096.
Najjar. RG: Pyke. CR; Adams. MB: Breitburg. D; Hershner. C: Kemp. M; Howarth. R;
Mulholland. MR: Paolisso. M; Secor. D; Sellner. K; Wardrop. D; Wood. R. (2010).
Potential climate-change impacts on the Chesapeake Bay. Estuar Coast Shelf Sci 86: 1-
20. http://dx.doi.Org/10.1016/j.ecss.2009.09.026.
Nanus. L; Clow. DW: Saros. JE; Stephens. VC: Campbell. DH. (2012). Mapping critical loads of
nitrogen deposition for aquatic ecosystems in the Rocky Mountains, USA. Environ Pollut
166: 125-135. http://dx.doi.Org/10.1016/j.envpol.2012.03.019.
Nanus. L; Mcmurrav. JA; Clow. DW: Saros. JE: Blett. T; Gurdak. JJ. (2017). Spatial variation of
atmospheric nitrogen deposition and critical loads for aquatic ecosystems in the Greater
Yellowstone Area. Environ Pollut 223: 644-656.
http://dx.doi.Org/10.1016/j.envpol.2017.01.077.
NAPAP (National Acid Precipitation Assessment Program). (1991). The U.S. National Acid
Precipitation Assessment Program 1990 Integrated Assessment Report. Washington, DC:
The NAPAP Office of the Director.
NAPAP (National Acid Precipitation Assessment Program). (2011). National acid precipitation
assessment program report to congress 2011: An integrated assessment. Washington, DC:
U.S. Environmental Protection Agency, Clean Air Markets Division.
http ://ny. water.usgs. gov/proi ects/NAPAP/.
June 2018
R-155
DRAFT: Do Not Cite or Quote

-------
NAPCA (National Air Pollution Control Administration). (1969). Air quality criteria for
particulate matter. Washington, DC.
Narita. D; Rehdanz. K; Tol. RSJ. (2012). Economic costs of ocean acidification: a look into the
impacts on global shellfish production. Clim Change 113: 1049-1063.
http://dx.doi.org/10.1007/slQ584-011-0383-3.
Nash. TH. Ill: Sigal, LL. (1999). Epiphytic lichens in the San Bernardino Mountains in relation
to oxidant gradients. In PR Miller; JR McBride (Eds.), Oxidant air pollution impacts in
the montane forests of southern California: A case study of the San Bernardino
Mountains (pp. 223-234). New York, NY: Springer, http://dx.doi.org/10.1007/978-1-
4612-1436-6 11.
Nasholm, T; Hogberg, P; Franklin. O; Metcalfe. D; Keel SG: Campbell C: Hurry. V: Linden S:
Hogberg, MN. (2013). Are ectomycorrhizal fungi alleviating or aggravating nitrogen
limitation of tree growth in boreal forests? New Phytol 198: 214-221.
http://dx.doi.org/10. Ill 1/nph. 12139.
Nasr. M; Castonguav. M; Ogilvie. J: Raymond. BA; Arp. PA. (2010). Modelling and mapping
critical loads and exceedances for the Georgia Basin, British Columbia, using a zero
base-cation depletion criterion. Journal of Limnology 69: 181-192.
http://dx.doi.org/10.3274/JL10-69-Sl-18.
Nave. LE; Vance. ED: Swanston. CW: Curtis. PS. (2009a). Impacts of elevated N inputs on
north temperate forest soil C storage, C/N, and net N-mineralization. Geoderma 153:
231-240. http://dx.doi.Org/10.1016/i.geoderma.2009.08.012.
Nave. LE: Vogel. CS: Gough. CM: Curtis. PS. (2009b). Contribution of atmospheric nitrogen
deposition to net primary productivity in a northern hardwood forest. Can J For Res 39:
1108-1118. http://dx.doi.org/10.1139/X09-038.
Navratil, T; Norton, SA; Fernandez, IJ; Nelson, SJ. (2010). Twenty-year inter-annual trends and
seasonal variations in precipitation and stream water chemistry at the Bear Brook
Watershed in Maine, USA. Environ Monit Assess 171: 23-45.
http://dx.doi.org/10.1007/slQ661-010-1527-z.
Neff, JC; Townsend, AR; Gleixner, G; Lehman, SJ: Turnbull, J: Bowman, WD. (2002). Variable
effects of nitrogen additions on the stability of turnover of soil carbon [Letter], Nature
419: 915-917. http://dx.doi.org/10.1038/natureO 1136.
Neff, KJ: Peyton, E; Shwartz, J: Henry, T; Robinson, RB. (2008). Episodic stream acidification
in the Great Smoky mountains national park: An investigation into the mechanisms of
acidification and impacts on native brook trout. In RW Babcock Jr; R Walton (Eds.),
World Environmental and Water Resources Congress 2008: AhupuaA (pp. 1-10).
Reston, VA: American Society of Civil Engineers.
http://dx.doi.org/10.1061/40976(316^)170.
Neff, KJ: Schwartz, JS: Henry, TB; Robinson, RB: Moore, SE; Kulp, MA. (2009). Physiological
stress in native southern brook trout during episodic stream acidification in the Great
Smoky Mountains National Park. Arch Environ Contam Toxicol 57: 366-376.
http://dx.doi.org/10.1007/s00244-0Q8-9269-4.
June 2018
R-156
DRAFT: Do Not Cite or Quote

-------
Neff. KJ; Schwartz. JS; Moore. SE; Kulp. MA. (2013). Influence of basin characteristics on
baseflow and stormflow chemistry in the Great Smoky Mountains National Park, USA.
Hydrolog Process 27: 2061-2074. http://dx.doi.ore/10.1002/hyp.9366.
Nelson. JL; Zavaleta. ES. (2012). Salt marsh as a coastal filter for the oceans: Changes in
function with experimental increases in nitrogen loading and sea-level rise. PLoS ONE 7:
e38558. http://dx.doi.org/10.1371/iournal.pone.0Q38558.
Nelson. SJ: Johnson. KB: Kahl. JS: Haines. TA; Fernandez. IJ. (2007). Mass balances of
mercury and nitrogen in burned and unburned forested watersheds at Acadia National
Park, Maine, USA. Environ Monit Assess 126: 69-80. http://dx.doi.org/10.1007/slQ661-
006-9332-4.
Nelson. WG. (2017). Development of an epiphyte indicator of nutrient enrichment: A critical
evaluation of observational and experimental studies. Ecol Indicat 79: 207-227.
http://dx.doi.Org/10.1016/i.ecolind.2017.04.034.
Nemergut. PR: Townsend. AR; Sattin. SR; Freeman. KR; Fierer. N: Neff. JC: Bowman. WD:
Schadt. CW: Weintraub. MN: Schmidt. SK. (2008). The effects of chronic nitrogen
fertilization on alpine tundra soil microbial communities: Implications for carbon and
nitrogen cycling. Environ Microbiol 10: 3093-3105. http://dx.doi.Org/10.l 11 l/j.1462-
2920.2008.01735.x.
Nemitz. E. (2000). Sources and sinks of ammonia within an oilseed rape canopy. Agr Forest
Meteorol 105: 385-404. http://dx.doi.org/10.1016/S0168-1923(W)00205-7.
Nemitz, E; Dorsey, JR; Flynn, MJ; Gallagher, MW: Hensen, A; Erisman, JW; Owen, SM;
Dammgen, U; Sutton, MA. (2009). Aerosol fluxes and particle growth above managed
grassland. Biogeosciences 6: 1627-1645. http://dx.doi.org/10.5194/bg-6-1627-2009.
Nemitz, E; Gallagher, MW: Duvzer, JH; Fowler, D. (2002). Micrometeorological measurements
of particle deposition velocities to moorland vegetation. Q J Roy Meteorol Soc 128:
2281-2300. http://dx.doi.org/10.1256/ai.01.71.
Nemitz, E; Sutton, MA: Wvers, GP; Oties, RP; Mennen, MG: van Putten, EM: Gallagher, MW.
(2004). Gas-particle interactions above a Dutch heathland: II. Concentrations and surface
exchange fluxes of atmospheric particles. Atmos Chem Phys 4: 1007-1024.
Nestlerode, JA; Hansen, YD: Teague, A: Harwell, MC. (2014). Application of a three-tier
framework to assess ecological condition of Gulf of Mexico coastal wetlands. Environ
Monit Assess 186: 3477-3493. http://dx.doi.org/10.1007/s 10661 -014-3631 -v.
Nevalainen, L; Ketola, M; Korosi, JB; Manca, M; Kurmaver, R; Koinig, KA; Psenner, R; Luoto,
TP. (2014). Zooplankton (Cladocera) species turnover and long-term decline of Daphnia
in two high mountain lakes in the Austrian Alps. Hydrobiologia 722: 75-91.
http://dx.doi.org/10.1007/slQ750-013-1676-5.
Newell, SE: McCarthy, MJ: Gardner, WS: Fulweiler, RW. (2016). Sediment Nitrogen Fixation: a
Call for Re-evaluating Coastal N Budgets. Estuaries Coasts 39: 1626-1638.
http://dx.doi. org/10.1007/s 1223 7-016-0116-v.
June 2018
R-157
DRAFT: Do Not Cite or Quote

-------
Newingham. BA; Vanier. CH; Charlet TN; Smith. SD. (2012). Effects of enhanced summer
monsoons, nitrogen deposition and soil disturbance on Larrea tridentata productivity and
subsequent herbivory in the Mojave Desert. J Arid Environ 87: 19-28.
http://dx.doi.Org/10.1016/i.iaridenv.2012.07.002.
Newman. JA; Abner. ML: Dado. RG: Gibson. DJ; Brookings. A: Parsons. AJ. (2003). Effects of
elevated C02, nitrogen and fungal endophyte-infection on tall fescue: Growth,
photosynthesis, chemical composition and digestibility. Global Change Biol 9: 425-437.
http://dx.doi.Org/10.1046/i.1365-2486.2003.00601.x.
Nezat. CA; Blum. JD; Driscoll. CT. (2010). Patterns of Ca/Sr and Sr-87/Sr-86 variation before
and after a whole watershed CaSi03 addition at the Hubbard Brook Experimental Forest,
USA. Geochim Cosmo Act 74: 3129-3142. http://dx.doi.Org/10.1016/i.gca.2010.03.013.
Nguyen. TB; Crounse. JD: Teng. AP; St. Clair. JM; Paulot. F; Wolfe. GM; Wennberg. PO.
(2015). Rapid deposition of oxidized biogenic compounds to a temperate forest. Proc
Natl Acad Sci USA 112: E392-E401. http://dx.doi.org/10.1073/pnas.1418702112.
Ni. X: Yang. W: Oi. Z: Liao. S: Xu. Z: Tan. B: Wang. B: Wu. O: Fu. C: You. C: Wu. F. (2017).
Simple additive simulation overestimates real influence: altered nitrogen and rainfall
modulate the effect of warming on soil carbon fluxes. Global Change Biol 23: 3371-
3381. http://dx.doi.org/10.Ill 1/gcb. 13588.
Nichols. BJ: Langdon. KR. (2007). The Smokies All Taxa biodiversity inventory: History and
progress. Southeastern Naturalist 6: 27-34.
Nickerson, MA; Krysko, KL; Owen, RD. (2002). Ecological status of the Hellbender
(Cryptobranchus alleganiensis) and the Mudpuppy (Necturus maculosus) salamanders in
the Great Smoky Mountains National Park. J N C Acad Sci 118: 27-34.
Nieboer. E; Richardson. DHS: Puckett. KJ; Tomassini. FD. (1976). The phytotoxicity of sulphur
dioxide in relation to measurable responses in lichens. In TA Mansfield (Ed.), Society for
Experimental Biology, Seminar Series: Volume 1, Effects of Air Pollutants on Plants.
Cambridge, United Kingdom: Cambridge University Press.
Nielsen. MG: Kahl. JS. (2007). Nutrient export from watersheds on Mt. Desert Island, Maine, as
a function of land use and fire history. Environ Monit Assess 126: 81-96.
http://dx.doi.org/10.1007/slQ661-006-9333-3.
Nierzwicki-Bauer. SA; Boylen. CW: Eichler. LW: Harrison. JP; Sutherland. JW: Shaw. W:
Daniels. RA; Charles. DF; Acker. FW: Sullivan. TJ; Momen. B; Bukaveckas. P. (2010).
Acidification in the Adirondacks: Defining the biota in trophic levels of 30 chemically
diverse acid-impacted lakes. Environ Sci Technol 44: 5721-5727.
http://dx.doi.org/10.1021/esl005626.
Nietsch. SL; Arnold. JG: Kiniry. JR; Srinivasan. R; Williams. JR. (2002). SWAT: Soil and water
assessment tool user's manual. (TWRI Report TR-192). College Station, TX: Texas
Water Resources Institute, USD A Agricultural Research Service.
http://swat.tamu.edu/media/1294/swatuserman.pdf.
June 2018
R-158
DRAFT: Do Not Cite or Quote

-------
Nilsen, TO; Ebbesson, LOE; Handeland, SO; Kroglund, F; Finstad, B; Angotzi, AR; Stefansson,
SO. (2013). Atlantic salmon (Salmo salar L.) smolts require more than two weeks to
recover from acidic water and aluminium exposure. Aquat Toxicol 142-143: 33-44.
http://dx.doi.Org/10.1016/i.aquatox.2013.07.016.
Nilsson. J; Grennfelt P. (1988). Critical loads for sulphur and nitrogen: Report from a workshop
held at Skokloster, Sweden, 19-24 March 1988. Skokloster, Sweden: Nordic Council of
Ministers: Copenhagen.
Nilsson. M, -C; Wardle, DA; Zackrisson, O; Jaderlund, A. (2002). Effects of alleviation of
ecological stresses on an alpine tundra community over an eight-year period. Oikos 97: 3-
17. http://dx.doi.Org/10.1034/i.1600-0706.2002.970101.x.
Niu, S; Classen. AT; Dukes. JS; Kardol, P; Liu. L; Luo, Y; Rustad, L; Sun. J; Tang. J; Templer,
PH; Thomas. RO; Tian. D; Vicca. S; Wang. YP; Xia. J; Zaehle. S. (2016). Global
patterns and substrate-based mechanisms of the terrestrial nitrogen cycle [Review], Ecol
Lett 19: 697-709. http://dx.doi.org/10.1111/ele. 12591.
Niu. Z; Zhang. X; Wang. S; Zeng. M; Wang. Z; Zhang. Y; Ci. Z. (2014). Field controlled
experiments on the physiological responses of maize (Zea mays L.) leaves to low-level
air and soil mercury exposures. Environ Sci Pollut Res Int 21: 1541-1547.
http://dx.doi.org/10.1007/sll356-013-2Q47-5.
Nixon. SW. (1995). Coastal marine eutrophication: A definition, social causes, and future
concerns. Ophelia41: 199-219. http://dx.doi.org/10.1080/00785236.1995.10422Q44.
Nixon, SW; Ammerman, JW; Atkinson, LP; Berounsky, VM; Billen, G; Boicourt, WC;
Boynton, WR; Church, TM; Ditoro, DM; Elmgren, R; Garber, JH; Giblin, AE; Jahnke,
RA; Owens, NJP; Pilson, MEQ; Seitzinger, SP. (1996). The fate of nitrogen and
phosphorus at the land sea margin of the North Atlantic Ocean. Biogeochemistry 35: 141-
180. http://dx.doi.org/10.1007/BF02179826.
NO A A (National Oceanic and Atmospheric Administration). (2017a). Coastwatch &
oceanwatch. Available online at https://coastwatch.noaa. gov/cw html/index.html
(accessed
NOAA (National Oceanic and Atmospheric Administration). (2017b). Gulf of Mexico harmful
algal bloom forecast. Available online at
https://tidesandcurrents.noaa.gov/hab/gomx.html (accessed
NOAA (National Oceanic and Atmospheric Administration). (2017c). Harmful algal bloom
observing system (HABSOS). Available online at https://habsos.noaa.gov/ (accessed
NOAA Fisheries Pacific Islands Regional Office. (2015a). Protected resources green sea turtles.
Available online at http://www.fpir.noaa.gov/PRD/prd green sea turtle.html (accessed
March 17, 2016).
NOAA Fisheries Pacific Islands Regional Office. (2015b). Protected resources green sea turtles:
Hawaiian green turtle life history [Fact Sheet], Honolulu, HI.
http://www.fpir.noaa.gov/Librarv/PRD/Sea%20Turtles/Fishing%20around%20sea%20tur
tles%20page/seaturtle-lifehistoryF AO. 2-4-11-1 ,pdf.
June 2018
R-159
DRAFT: Do Not Cite or Quote

-------
Nob re. AM; Ferreira. JG; Newton. A; Si mas. T; Icelv. JD; Neves. R. (2005). Managing
eutrophication: Integration of field data, ecosystem-scale simulations and screening
models. J Mar Syst 56: 375-390. http://dx.doi.ore/10.1016/i.imarsys.2005.03.003.
Nodine. ER; Gaiser. EE. (2014). Distribution of diatoms along environmental gradients in the
Charlotte Harbor, Florida (USA), estuary and its watershed: Implications for
bioassessment of salinity and nutrient concentrations. Estuaries Coasts 37: 864-879.
http://dx.doi. org/10.1007/s 1223 7-013 -9729-6.
Nogueira. MA: Magalhaes. GC: Cardoso. EJB. N. (2004). Manganese toxicity in mycorrhizal
and phosphorus-fertilized soybean plants. J Plant Nutr 27: 141-156.
http://dx.doi.org/10.1081/PLN-120027552.
Noh. S: Kim. CK; Lee. JH; Kim. Y; Choi. K; Han. S. (2016). Physicochemical factors affecting
the spatial variance of monomethylmercury in artificial reservoirs. Environ Pollut 208:
345-353. http://dx.doi.Org/10.1016/i.envpol.2015.09.054.
Nolte. CG: Appel. KW: Kelly. JT; Bhave. PV: Fahev. KM: Collett. JL. Jr; Zhang. L; Young. JO.
(2015). Evaluation of the Community Multiscale Air Quality (CMAQ) model v5.0
against size-resolved measurements of inorganic particle composition across sites in
North America. Geoscientific Model Development 8: 2877-2892.
http://dx.doi.org/10.5194/gmd-8-2877-2015.
Norbv. RJ; Weerasuriya. Y; Hanson. PJ. (1989). Induction of nitrate reductase activity in red
spruce needles by N02 and HN03 vapor. Can J For Res 19: 889-896.
http://dx.doi.Org/10.l 139/x89-135.
Norton. S: Kahl. J: Fernandez. I; Haines. T; Rustad. L; Nodvin. S: Scofield. J: Strickland. T;
Erickson. J: Wigington. P. Jr: Lee. J. (1999). The Bear Brook Watershed, Maine
(BBWM), USA. Environ Monit Assess 55: 7-51.
http://dx.doi.Org/10.1023/A:1006115011381.
Norton. SA; Fernandez. IJ; Kahl. JS: Reinhardt. RL. (2004). Acidification trends and the
evolution of neutralization mechanisms through time at the Bear Brook Watershed in
Maine (BBWM), U.S.A. In Biogeochemical Investigations of Terrestrial, Freshwater, and
Wetland Ecosystems Across the Globe. New York, NY: Springer.
http://dx.doi.org/10.1007/978-94-007-Q952-2 21.
Norton. SA: Fernandez. IJ: Kahl. JS: Rustad. LE; Navratil. T; Almquist. H. (2010). The
evolution of the science of Bear Brook Watershed in Maine, USA. Environ Monit Assess
171: 3-21. http ://dx. doi .org/10.1007/s 10661 -010-1528-v.
Nouchi. I: Mavumi. H; Yamazoe. F. (1984). Foliar injury response of petunia and kidney bean to
simultaneous and alternate exposures to ozone and PAN. Atmos Environ 18: 453-460.
http ://dx. doi .org/10.1016/0004-6981 (84^)90122-7.
Nowicki. BL: Nixon. SW. (1985). BENTHIC NUTRIENT REMINERALIZATION IN A
COASTAL LAGOON ECOSYSTEM. Estuaries 8: 182-190.
Nowlan. CR; Martin. RV; Philip. S: Lamsal. LN: Krotkov. NA; Marais. EA; Wang. S: Zhang. O.
(2014). Global dry deposition of nitrogen dioxide and sulfur dioxide inferred from space-
based measurements. Global Biogeochem Cycles 28: 1025-1043.
http://dx.doi.org/10.1002/2014GB0048Q5.
June 2018
R-160
DRAFT: Do Not Cite or Quote

-------
NPS (U.S. National Park Service). (2013). Species list for Rocky Mountain National Park.
NPSpecies database [Database], Washington, DC. Retrieved from
https://irma.nps.gov/NPSpecies/Search/SpeciesList/ROMQ
NRC (National Research Council). (2000). Clean coastal waters: Understanding and reducing the
effects of nutrient pollution. In Clean Coastal Waters: Understanding and Reducing the
Effects of Nutrient Pollution. Washington, DC: The National Academies Press.
http://dx.doi.ore/10.17226/9812.
NRC (National Research Council). (2004). Air quality management in the United States.
Washington, DC: National Academies Press.
http://www.nap.edu/catalog.php7record id=10728.
NRCS (Natural Resources Conservation Service). (2009). PLANTS database. Available online at
http://plants.usda.gov/iava/ (accessed
Nussbaum, S; von Ballmoos, P; Gfeller, H; Schlunegger, UP; Fuhrer, J; Rhodes, D; Brunold, C.
(1993). Incorporation of atmospheric 15N02-nitrogen into free amino acids by Norway
spruce Picea abies (L.) Karst. Oecologia 94: 408-414.
http://dx.doi.org/10.1007/BF0Q317117.
Nybakken, L; Johansson, O; Palmqvist K. (2009). Defensive compound concentration in boreal
lichens in response to simulated nitrogen deposition. Global Change Biol 15: 2247-2260.
http://dx.doi.org/10.111 l/i.l365-2486.2009.01853.x.
Nvdick, KR; Lafrancois, BM; Baron, JS: Johnson, BM. (2004). Nitrogen regulation of algal
biomass, productivity, and composition in shallow mountain lakes, Snowy Range,
Wyoming, USA. Can J Fish Aquat Sci 61: 1256-1268. http://dx.doi.org/10.1139/F04-
085.
O'Dea, CB; Anderson, S: Sullivan, T; Landers, D; Casey, CF. (2017). Impacts to ecosystem
services from aquatic acidification: using FEGS-CS to understand the impacts of air
pollution. Ecosphere 8. http://dx.doi.org/10.1002/ecs2.1807.
O'Driscoll, MA: DeWalle, DR. (2010). Seeps regulate stream nitrate concentration in a forested
Appalachian catchment. J Environ Qual 39: 420-431.
http://dx.doi.org/10.2134/ieq2009.0Q83.
O'Toole, P; Robson, BJ; Chambers, JM. (2017). Riparian vegetation condition is associated with
invertebrate assemblage composition in intermittent and humic streams. Aquat Sci 79:
277-289. http://dx.doi.org/10.1007/s00027-016-Q496-x.
OAOPS-Emissions Inventory and Analysis Group (Office of Air Quality Planning and Standards
Emissions Inventory and Analysis Group). (2016). Profile of version 1 of the 2014
National Emissions Inventory. Washington, D.C.: U.S. Environmental Protection
Agency.
Ochoa-Hueso, R; Delgado-Baquerizo, M; Gallardo, A: Bowker, MA: Maestre, FT. (2016).
Climatic conditions, soil fertility and atmospheric nitrogen deposition largely determine
the structure and functioning of microbial communities in biocrust-dominated
Mediterranean drylands. Plant Soil 399: 271-282. http://dx.doi.org/10.1007/sl 1104-015-
2695-y.
June 2018
R-161
DRAFT: Do Not Cite or Quote

-------
Ochoa-Hueso. R; Manrique. E. (2010). Nitrogen fertilization and water supply affect
germination and plant establishment of the soil seed bank present in a semi-arid
Mediterranean scrubland. Plant Ecol 210: 263-273. http://dx.doi.ore/10.1007/sl 1258-
010-9755-4.
Ochoa-Hueso. R; Rocha, I: Stevens. CJ: Manrique, E; Jose Lucianez, M. (2014). Simulated
nitrogen deposition affects soil fauna from a semiarid Mediterranean ecosystem in central
Spain. Biol Fertil Soils 50: 191-196. http://dx.doi.org/10.1007/sQ0374-013-0838-v.
Ochoa-Hueso. R; Stevens. CJ. (2015). European semiarid Mediterranean ecosystems are
sensitive to nitrogen deposition: Impacts on plant communities and root phosphatase
activity. Water Air Soil Pollut 226. http://dx.doi.org/10.1007/sl 1270-014-2278-1.
Oczkowski. A: Wigand. C: Hanson. A: Markham. E; Miller. KM: Johnson. R. (2016). Nitrogen
Retention in Salt Marsh Systems Across Nutrient-Enrichment, Elevation, and
Precipitation Regimes: a Multiple-Stressor Experiment. Estuaries Coasts 39: 68-81.
http://dx.doi.org/10.1007/sl2237-015-9975-x.
Office of Air and Radiation (EPA Office of Air and Radiation). (2011). The benefits and costs of
the Clean Air Act from 1990 to 2020. Final report - Rev. A [EPA Report], (EPA 402-R-
11-004).	WAshington, D.C.: U.S. Environmental Protection Agency.
Officer. CB; Biggs. RB; Taft. JL; Cronin. LE; Tyler. MA: Boynton. WR. (1984). Chesapeake
Bay anoxia: Origin, development, and significance. Science 223: 22-27.
http://dx.doi.Org/10.l 126/science.223.4631.22.
Oguntimehin, I: Eissa, F; Sakugawa, H. (2010). Negative effects of fluoranthene on the
ecophysiology of tomato plants (Lycopersicon esculentum Mill) Fluoranthene mists
negatively affected tomato plants. Chemosphere 78: 877-884.
http://dx.doi. org/10.1016/i. chemosphere.2009.11.030.
Oguntimehin, I: Nakatani, N; Sakugawa, H. (2008). Phytotoxicities of fluoranthene and
phenanthrene deposited on needle surfaces of the evergreen conifer, Japanese red pine
(Pinus densiflora Sieb. etZucc.). Environ Pollut 154: 264-271.
http://dx.doi.Org/10.1016/i.envpol.2007.10.039.
Ohta, T; Niwa, S; Agetsuma, N; Hiura, T. (2014). Calcium concentration in leaf litter alters the
community composition of soil invertebrates in warm-temperate forests. Pedobiologia 57:
257-262. http://dx.doi.Org/10.1016/i.pedobi.2014.07.003.
Oikawa, PY; Ge, C: Wang, J: Eberwein, JR; Liang, LL; Allsman, LA: Grantz, DA: Jenerette,
GD. (2015). Unusually high soil nitrogen oxide emissions influence air quality in a high-
temperature agricultural region. Nat Commun 6: 8753.
http://dx.doi.org/10.1038/ncomms9753.
Oia, T; Arp, PA. (1998). Assessing atmospheric sulfur and nitrogen loads critical to the
maintenance of upland forest soils. In DG Maynard (Ed.), Soils, Plants, and the
Environment (pp. 337-363). New York, NY: Marcel Dekker.
Olivares, A: van Drooge, BL; Casado, M; Prats, E, va: Serra, M; van Per Yen, L; Kamstra, JH;
Hamers, T; Hermsen, S: Grimalt, JO: Pina, B. (2013). Developmental effects of aerosols
and coal burning particles in zebrafish embryos. Environ Pollut 178: 72-79.
http://dx.doi.Org/10.1016/i.envpol.2013.02.026.
June 2018
R-162
DRAFT: Do Not Cite or Quote

-------
Oliver. SK; Collins. SM; Soranno. PA; Wagner. T; Stanley. EH; Jones. JR; Stow. CA; Lottig.
NR. (2017). Unexpected stasis in a changing world: Lake nutrient and chlorophyll trends
since 1990. Global Change Biol 23: 5455-5467. http://dx.doi.org/10. Ill 1/gcb. 13810.
Olyarnik. SV; Stachowicz. JJ. (2012). Multi-year study of the effects of Ulva sp blooms on
eelgrass Zostera marina. Mar Ecol Prog Ser 468: 107-117.
http://dx.doi.org/10.3354/meps09973.
Omernik. JM; Powers. CF. (1983). Map supplement: Total alkalinity of surface waters-a national
map. Ann Assoc Am Geogr 73: 133-136. http://dx.doi.Org/10.l 11 l/j.1467-
8306.1983.tb01400.x.
Onipchenko. VG; Makarov. MI; Akhmetzhanova. AA; Soudzilovskaia. NA; Aibazova. FU;
Elkanova. MK; Stogova. AY; Cornelissen. JHC. (2012). Alpine plant functional group
responses to fertiliser addition depend on abiotic regime and community composition.
Plant Soil 357: 103-115. http://dx.doi.org/10.1007/sl 1104-012-1146-2.
Orem. W; Gilmour. C; Axelrad. D; Krabbenhoft. D; Scheidt. D; Kalla. P; Mccormick. P;
Gabriel. M; Aiken. G. (2011). Sulfur in the South Florida ecosystem: Distribution,
sources, biogeochemistry, impacts, and management for restoration. Crit Rev Environ Sci
Tech 41: 249-288. http://dx.doi.org/10.1080/10643389.201Q.531201.
Ormerod. SJ; Boole. P; McCahon. CP; Weatherlev. NS; Pascoe. D; Edwards. RW. (1987). Short-
term experimental acidification of a Welsh stream: Comparing the biological effects of
hydrogen ions and aluminum. Freshw Biol 17: 341-356. http://dx.doi.Org/10.l 11 l/j.1365-
2427.1987.tb01054.x.
Ormrod. DP. (1984). Impact of trace element pollution on plants. In M Treshow (Ed.), Air
pollution and plant life (pp. 291-314). Chichester, United Kingdom: John Wiley & Sons
Ltd.
Orr, JC; Fabry, VJ; Aumont, O; Bopp, L; Doney, SC; Feely, RA; Gnanadesikan, A; Gruber, N;
Ishida. A; Joos. F; Key. RM; Lindsay. K; Maier-Reimer. E; Matear. R; Monfrav. P;
Mouchet. A; Najjar. RG; Plattner. GK; Rodgers. KB; Sabine. CL; Sarmiento. JL;
Schlitzer. R; Slater. RD; Totterdell. U; Weirig. MF; Yamanaka. Y; Yool. A. (2005).
Anthropogenic ocean acidification over the twenty-first century and its impact on
calcifying organisms. Nature 437: 681-686. http://dx.doi.org/10.1038/nature04095.
Orth. RJ; Carruthers. TJB; Dennison. WC; Duarte. CM; Fourqurean. JW; Heck. KL. Jr; Hughes.
AR; Kendrick. GA; Kenworthy. WJ; Olyarnik. S; Short. FT; Wavcott. M; Williams. SL.
(2006). A global crisis for seagrass ecosystems. Bioscience 56: 987-996.
Orth. RJ; Williams. MR; Marion. SR; Wilcox. DJ; Carruthers. TJB; Moore. KA; Kemp. WM;
Dennison. WC; Rybicki. N; Bergstrom. P; Batiuk. RA. (2010). Long-term trends in
submersed aquatic vegetation (SAV) in Chesapeake Bay, USA, related to water quality.
Estuaries Coasts 33: 1144-1163. http://dx.doi.org/10.1007/sl2237-01Q-9311-4.
Ortiz-Montalvo. PL; Hakkinen. SAK; Schwier. AN; Lim. YB; McNeill. VF; Turpin. BJ. (2014).
Ammonium addition (and aerosol pH) has a dramatic impact on the volatility and yield of
glyoxal secondary organic aerosol. Environ Sci Technol 48: 255-262.
http://dx.doi.org/10.1021/es4035667.
June 2018
R-163
DRAFT: Do Not Cite or Quote

-------
Ostman. O; Ekl of. J; Eriksson. BK; Olsson. J; Moksnes. P. erO; Bergstrom. U. If. (2016). Top-
down control as important as nutrient enrichment for eutrophication effects in North
Atlantic coastal ecosystems. J Appl Ecol 53: 1138-1147. http://dx.doi.Org/10.l 111/1365-
2664.12654.
Ou. M; Hamilton. TJ; Eom. J: Lyall. EM: Gallup. J: Jiang. A: Lee. J: Close. DA: Yun. S. -S:
Brauner. CJ. (2015). Responses of pink salmon to C02-induced aquatic acidification. Nat
Clim Chang 5: 950-955. http://dx.doi.org/10.1038/NCLIMATE2694.
Ouimet. R: Arp. PA: Watmough. SA: Aherne. J: DeMerchant. I. (2006). Determination and
mapping critical loads of acidity and exceedances for upland forest soils in eastern
Canada. Water Air Soil Pollut 172: 57-66. http://dx.doi.org/10.1007/sll270-005-905Q-5.
Ouimet. R: Duchesne. L: Houle. D: Arp. PA. (2001). Critical loads and exceedances of acid
deposition and associated forest growth in the northern hardwood and boreal coniferous
forests in Quebec, Canada. Water Air Soil Pollut 1: 119-134.
http://dx.doi.Org/10.1023/A:1011544325004.
Oulehle. F; Cosby. BJ; Wright. RF; Hruska. J: Kopacek. J: Kram. P; Evans. CD: Moldan. F.
(2012). Modelling soil nitrogen: The MAGIC model with nitrogen retention linked to
carbon turnover using decomposer dynamics. Environ Pollut 165: 158-166.
http://dx.doi.Org/10.1016/i.envpol.2012.02.021.
Ow. YX; Vogel. N: Collier. CJ: Holtum. JAM: Flores. F; Uthicke. S. (2016). Nitrate fertilisation
does not enhance C02 responses in two tropical seagrass species. Sci Rep 6: 23093.
http://dx.doi. org/10.1038/srep23 093.
Ozen. A: Sorf. M; Trochine. C: Liboriussen. L; Beklioglu. M; S0ndergaard. M; Lauridsen. TL:
Johansson. LS: Jeppesen. E. (2013). Long-term effects of warming and nutrients on
microbes and other plankton in mesocosms. Freshw Biol 58: 483-493.
http://dx.doi.org/10.111 l/i,1365-2427.2012.02824.x.
Pabian. SE: Brittingham. MC. (2012). Soil calcium and forest birds: Indirect links between
nutrient availability and community composition. Ecosystems 15: 748-760.
http://dx.doi.org/10.1007/slQ021-012-9543-l.
Paciga, AL; Riipinen, I; Pandis, SN. (2014). Effect of ammonia on the volatility of organic
diacids. Environ Sci Technol 48: 13769-13775. http://dx.doi.org/10.1021/es5037805.
Padgett. P; Allen. EB. (1999). Differential responses to nitrogen fertilization in native shrubs and
exotic annuals common to Mediterranean coastal sage scrub of California. Plant Ecol
144: 93-101. http://dx.doi.Org/10.1023/A:1009895720067.
Padgett. PE: Allen. EB: Bytnerowicz. A: Minich. RA. (1999). Changes in soil inorganic nitrogen
as related to atmospheric nitrogenous pollutants in southern California. Atmos Environ
33: 769-781. http://dx.doi.org/10.1016/S1352-2310(98V)0214-3.
Padgett. PE: Cook. H: Bytnerowicz. A: Heath. RL. (2009a). Foliar loading and metabolic
assimilation of dry deposited nitric acid air pollutants by trees. J Environ Monit 11: 75-
84. http://dx.doi.org/10.1039/b8Q4338h.
June 2018
R-164
DRAFT: Do Not Cite or Quote

-------
Padgett. PE; Parry. SD; Bytnerowicz. A; Heath. RL. (2009b). Image analysis of epicuticular
damage to foliage caused by dry deposition of the air pollutant nitric acid. J Environ
Monit 11: 63-74. http://dx.doi.org/10.1039/b8Q4875d.
Padmavathiamma. PK; Li. LY. (2007). Phytoremediation technology: Hyper-accumulation
metals in plants. Water Air Soil Pollut 184: 105-126. http://dx.doi.org/10.1007/sll27Q-
007-9401-5.
Paerl. HW. (1997). Coastal eutrophication and harmful algal blooms: Importance of atmospheric
deposition and groundwater as "new" nitrogen and other nutrient sources. Limnology 42:
1154-1162. http://dx.doi.Org/10.4319/lo.1997.42.5 part 2.1154.
Paerl. HW. (2009). Controlling eutrophi cation along the freshwater-marine continuum: Dual
nutrient (N and P) reductions are essential. Estuaries Coasts 32: 593-601.
http://dx.doi.org/10.1007/sl2237-0Q9-9158-8.
Paerl, HW; Boynton, WR; Dennis, RL; Driscoll, CT; Greening, HS; Kremer, JN; Rabalais, NN;
Seitzinger, SP. (2000). Atmospheric deposition of nitrogen in coastal waters:
Biogeochemical and ecological implications. In RW Valigura; RB Alexander; MS
Castro; TP Meyers; HW Paerl; PE Stacey; RE Turner (Eds.), Nitrogen Loading in
Coastal Water Bodies: An Atmospheric Perspective (pp. 11-52). Washington, DC:
American Geophysical Union. http://dx.doi.org/10.1029/CE057pQ011.
Paerl, HW; Dennis, RL; Whitall, DR. (2002). Atmospheric deposition of nitrogen: Implications
for nutrient over-enrichment of coastal waters. Estuaries Coasts 25: 677-693.
http://dx.doi.org/10.1007/BF028Q4899.
Paerl, HW; Gardner, WS; Havens, KE; Joyner, AR; McCarthy, MJ; Newell, SE; Qin, B; Scott,
JT. (2016a). Mitigating cyanobacterial harmful algal blooms in aquatic ecosystems
impacted by climate change and anthropogenic nutrients [Review], Harmful Algae 54:
213-222. http://dx.doi.Org/10.1016/i.hal.2015.09.009.
Paerl, HW; Hall, NS; Calandrino, ES. (2011). Controlling harmful cyanobacterial blooms in a
world experiencing anthropogenic and climatic-induced change [Review], Sci Total
Environ 409: 1739-1745. http://dx.doi.Org/10.1016/i.scitotenv.2011.02.001.
Paerl, HW; Hall, NS; Peierls, BL; Rossignol, KL. (2014). Evolving paradigms and challenges in
estuarine and coastal eutrophi cation dynamics in a culturally and climatically stressed
world. Estuaries Coasts 37: 243-258. http://dx.doi.org/10.1007/sl2237-014-9773-x.
Paerl, HW; Justic, D. (2011). Primary producers: Phytoplankton ecology and trophic dynamics in
coastal waters. In Treatise on Estuarine and Coastal Science Primary Producers. London,
England: Academic Press. http://dx.doi.org/10.1016/B978-Q-12-374711-2.00603-3.
Paerl, HW; Otten, TG. (2013). Harmful cyanobacterial blooms: Causes, consequences, and
controls. Microb Ecol 65: 995-1010. http://dx.doi.org/10.1007/sQ0248-012-0159-v.
Paerl, HW; Piehler, MF. (2008). Nitrogen and marine eutrophication. In DG Capone; DA Bronk;
MR Mulholland; EJ Carpenter (Eds.), Nitrogen in the Marine Environment (2nd ed., pp.
529-567). San Diego, CA: Academic Press. http://dx.doi.org/10.1016/B978-Q-12-372522-
6.00011-6.
June 2018
R-165
DRAFT: Do Not Cite or Quote

-------
Paerl. HW; Rossignol. KL; Hall. SN; Peierls. BL; Wetz. MS. (2010). Phytoplankton community
indicators of short- and long-term ecological change in the anthropogenically and
climatically impacted Neuse River estuary, North Carolina, USA. Estuaries Coasts 33:
485-497. http://dx.doi.org/10.1007/sl2237-009-9137-Q.
Paerl HW: Scott JT. (2010). Throwing fuel on the fire: Synergistic effects of excessive nitrogen
inputs and global warming on harmful algal blooms. Environ Sci Technol 44: 7756-7758.
http://dx.doi.ore/10.1021/esl02665e.
Paerl HW: Scott JT: McCarthy. MJ; Newell SE; Gardner. WS: Havens. KE; Hoffman. DK;
Wilhelm. SW: Wurtsbaugh. WA. (2016b). It Takes Two to Tango: When and Where
Dual Nutrient (N & P) Reductions Are Needed to Protect Lakes and Downstream
Ecosystems. Environ Sci Technol. http://dx.doi.org/10.1021/acs.est.6b02575.
Paerl. HW: Valdes. LM; Pincknev. JL; Piehler. MF; Dyble. J: Moisander. PH. (2003).
Phytoplankton photopigments as indicators of estuarine and coastal eutrophication.
Bioscience 53: 953-964.
Paerl. HW: Whitall. DR. (1999). Anthropogenically-derived atmospheric nitrogen deposition,
marine eutrophication and harmful algal bloom expansion: Is there a link? Ambio 28:
307-311.
Page. BP: Mitchell. MJ. (2008). Influences of a calcium gradient on soil inorganic nitrogen in
the Adirondack Mountains, New York. Ecol Appl 18: 1604-1614.
http://dx.doi.Org/10.1890/07-0150.l.
Pak, K, -R; Bartha, R. (1998). Mercury methylation and demethylation in anoxic lake sediments
and by strictly anaerobic bacteria. Appl Environ Microbiol 64: 1013-1017.
Palacios. SL; Zimmerman. RC. (2007). Response of eelgrass Zostera marina to C02 enrichment:
possible impacts of climate change and potential for remediation of coastal habitats. Mar
Ecol Prog Ser 344: 1-13. http://dx.doi.org/10.3354/meps07084.
Palmqvist. K. (2000). Carbon economy in lichens. New Phytol 148: 11-36.
http://dx.doi.Org/10.1046/i.1469-8137.2000.00732.x.
Pandev. J: Pandev. U. (2009). Accumulation of heavy metals in dietary vegetables and cultivated
soil horizon in organic farming system in relation to atmospheric deposition in a
seasonally dry tropical region of India. Environ Monit Assess 148: 61-74.
http://dx.doi.org/10.1007/slQ661-007-0139-8.
Pannatier. EG: Thimonier. A: Schmitt. M; Walthert. L; Waldner. P. (2011). A decade of
monitoring at Swiss Long-Term Forest Ecosystem Research (LWF) sites: Can we
observe trends in atmospheric acid deposition and in soil solution acidity. Environ Monit
Assess 174: 3-30. http://dx.doi.org/10.1007/slQ661-010-1754-3.
Pannek. A: Dupre. C: Gowing. DJG: Stevens. CJ: Diekmann. M. (2015). Spatial gradient in
nitrogen deposition affects plant species frequency in acidic grasslands. Oecologia 177:
39-51. http://dx.doi.org/10.1007/sQ0442-014-3120-6.
June 2018
R-166
DRAFT: Do Not Cite or Quote

-------
Paoli, L; Guttova, A; Grassi, A; Lackovicova, A; Senko, D; Sorbo, S; Basile, A; Loppi, S.
(2015). Ecophysiological and ultrastructural effects of dust pollution in lichens exposed
around a cement plant (SW Slovakia). Environ Sci Pollut Res Int 22: 15891-15902.
http://dx.doi.org/10.1007/sll356-015-48Q7-x.
Paranjape. AR; Hall BP: Smol. J. (2017). Recent advances in the study of mercury methylation
in aquatic systems. 2: 85-119.
Pardo. LH; Duarte. N. (2007). Assessment of effects of acidic deposition on forested ecosystems
in Great Smoky Mountains National Park using critical loads for sulfur and nitrogen. S.
Burlington, VT: USDA Forest Service.
http://www.nrs.fs.fed.us/clean air water/clean water/critical loads/local-
resources/docs/GSMNP CL Report 070830.pdf.
Pardo. LH: Fenn, ME: Goodale, CL; Geiser, LH: Priscoll, CT; Allen. EB; Baron. JS: Bobbink,
R: Bowman. WD: Clark. CM: Emmett. B: Gilliam. FS: Greaver. TL: Hall. SJ: Lilleskov.
EA: Liu. L: Lynch. JA: Nadelhoffer. KJ: Perakis. SS: Robin-Abbott. MJ: Stoddard. JL:
Weathers, KC: Dennis, RL. (201 la). Effects of nitrogen deposition and empirical
nitrogen critical loads for ecoregions of the United States. Ecol Appl 21: 3049-3082.
http://dx.doi.Org/10.1890/10-2341.l.
Pardo, LH: Geiser, LH: Fenn, ME: Driscoll, CT: Goodale, CL: Allen, EB: Baron, JS: Bobbink,
R: Bowman, WD: Clark, CM: Emmett, B: Gilliam, FS: Greaver, TL: Hall, SJ: Lilleskov,
EA: Liu, L: Lynch, JA: Nadelhoffer, K: Perakis, SS: Robin-Abbott, MJ: Stoddard, JL:
Weathers, KC. (201 lb). Synthesis. In Assessment of nitrogen deposition effects and
empirical critical loads of nitrogen for ecoregions of the United States (General Technical
Report NRS-80) (pp. 229-284). (NRS-80). Newtown Square, PA: U.S. Department of
Agriculture, Forest Service, Northern Research Station.
http://www.nrs.fs.fed.us/pubs/gtr/gtr-nrs-80chapters/19-pardo.pdf.
Pardo, LH: Robin-Abbott, MJ: Driscoll, CT. (201 lc). Assessment of nitrogen deposition effects
and empirical critical loads of nitrogen for ecoregions of the United States. In Assessment
of nitrogen deposition effects and empirical critical loads of Nitrogen for ecoregions of
the United States. (NRS-80). Newtown Square, PA: U.S. Department of Agriculture,
Forest Service, Northern Research Station, http://www.nrs.fs.fed.us/pubs/38109.
Parkhurst, PL: Appelo, CAJ. (1999). UseriEs guide to PHREEQC (version 2)—A computer
program for speciation, batch-reaction, one-dimensional transport, and inverse
geochemical calculations. (99-4259). Denver, CO: U.S. Geological Survey.
Parrent, JL: Vilgalys, R. (2009). Expression of genes involved in symbiotic carbon and nitrogen
transport in Pinus taeda mycorrhizal roots exposed to C02 enrichment and nitrogen
fertilization. Mycorrhiza 19: 469-479. http://dx.doi.org/10.1007/s00572-009-025Q-5.
Parrish, ZD: White, JC: Isleven, M: Gent, MPN: Iannucci-Berger, W: Eitzer, BP: Kelsev, JW:
Mattina, MI. (2006). Accumulation of weathered polycyclic aromatic hydrocarbons
(PAHs) by plant and earthworm species. Chemosphere 64: 609-618.
http://dx.doi.Org/10.1016/i.chemosphere.2005.l 1.003.
Parsons, BG: Watmough, SA: Pillon, PJ: Somers, KM. (2010). A bioassessment of lakes in the
Athabasca Oil Sands Region, Alberta, using benthic macroinvertebrates. Journal of
Limnology 69: 105-117. http://dx.doi.org/10.3274/JL10-69-Sl-ll.
June 2018
R-167
DRAFT: Do Not Cite or Quote

-------
Parton. WJ; Hart man. M; Oiima. D; Schimel. D. (1998). DAYCENT and its land surface
submodel: description and testing. Glob Planet Change 19: 35-48.
Pasari. JR; Hernandez. PL: Zavaleta. ES. (2014). Interactive effects of nitrogen deposition and
grazing on plant species composition in a serpentine grassland. Rangeland Ecol Manag
67: 693-700. http://dx.doi.org/10.2111/REM-D-13-00116.1.
Pasquini. SC: Vourlitis. GL. (2010). Post-fire primary production and plant community
dynamics in chaparral stands exposed to varying levels of nitrogen deposition. J Arid
Environ 74: 310-314. http://dx.doi.Org/10.1016/i.iaridenv.2009.08.013.
Passy, SI. (2008). Continental diatom biodiversity, in stream benthos declines as more nutrients
become limiting. ProcNatl Acad Sci USA 105: 9663-9667.
http://dx.doi.org/10.1073/pnas.08025421Q5.
Pastor. J: Dewey. B; Johnson. NW; Swain. EB; Monson. P; Peters. EB; Myrbo. A. (2017).
Effects of sulfate and sulfide on the life cycle of Zizania palustris in hydroponic and
mesocosm experiments. Ecol Appl 27: 321-336. http://dx.doi.org/10.1002/eap.1452.
Pastore. MA: Megonigal. JP; Langlev. JA. (2016). Elevated C02 promotes long-term nitrogen
accumulation only in combination with nitrogen addition. Global Change Biol 22: 391-
403. http://dx.doi.org/10. Ill 1/gcb. 13112.
Pataki. DE; Carreiro. MM: Cherrier. J: Grulke. NE; Jennings. V: Pincetl. S: Pouvat. RV:
Whitlow. TH; Zipperer. WC. (2011). Coupling biogeochemical cycles in urban
environments: Ecosystem services, green solutions, and misconceptions. Front Ecol
Environ 9: 27-36. http://dx.doi.org/10.1890/09022Q.
Patra. M; Bhowmik. N: Bandopadhyav. B; Sharma. A. (2004). Comparison of mercury, lead and
arsenic with respect to genotoxic effects on plant systems and the development of genetic
tolerance. Environ Exp Bot 52: 199-223.
http://dx.doi.Org/10.1016/i.envexpbot.2004.02.009.
Patterson. SL; Zak. PR: Burton. AJ; Talhelm. AF; Pregitzer. KS. (2012). Simulated N deposition
negatively impacts sugar maple regeneration in a northern hardwood ecosystem. J Appl
Ecol 49: 155-163. http://dx.doi.org/10.1111/i. 1365-2664.2011,02090.x.
Paudel. B; Montagna. PA: Besonen. M; Adams. L. (2017). Inorganic nitrogen release from
sediment slurry of riverine and estuarine ecosystems located at different river regimes.
Mar Freshwat Res 68: 1282-1291. http://dx.doi.org/10.1071/MF1626Q.
Paul. AJ: Leavitt. PR: Schindler. DW: Hardie. AK. (1995). Direct and indirect effects of
predation by a calanoid copepod (subgenus: Hesperodiaptomus) and of nutrients in a
Ashless alpine lake. Can J Fish Aquat Sci 52: 2628-2638. http://dx.doi.org/10.1139/f95-
852.
Pauli. D; Peintinger. M; Schmid. B. (2002). Nutrient enrichment in calcareous fens: Effects on
plant species and community structure. Basic Appl Ecol 3: 255-266.
http://dx.doi.org/10.1078/1439-1791-00Q96.
Paulissen, MPC, P; Bobbink, R; Robat, SA; Verhoeven, J, osTA. (2016). Effects of Reduced and
Oxidised Nitrogen on Rich-Fen Mosses: a 4-Year Field Experiment. Water Air Soil
Pollut 227. http://dx.doi.org/10.1007/sll270-Q15-2713-v.
June 2018
R-168
DRAFT: Do Not Cite or Quote

-------
Paulo. LM; Stains. AJM; Sousa. DZ. (2015). Methanogens, sulphate and heavy metals: A
complex system. Reviews in Environmental Science and Biotechnology 14: 537-553.
http://dx.doi.org/10.1007/slll57-Q15-9387-l.
Paulot F; Crounse. JD; Kiaergaard. HG: Kroll. JH: Seinfeld. JH: Wennberg. PO. (2009).
Isoprene photooxidation: New insights into the production of acids and organic nitrates.
Atmos Chem Phys 9: 1479-1501. http://dx.doi.org/10.5194/acp-9-1479-2009.
Paulot F; Jacob. DJ: Pinder. RW: Bash. JO: Travis. K; Henze. DK. (2014). Ammonia emissions
in the United States, European Union, and China derived by high-resolution inversion of
ammonium wet deposition data: Interpretation with a new agricultural emissions
inventory (MASAGE_NH3). J Geophys Res Atmos 119: 4343-4364.
http://dx.doi.org/10.1002/2013JD02113Q.
Paulot F; Wunch. D; Crounse. JD: Toon. GC: Millet DB; DeCarlo. PF; Vigouroux. C:
Deutscher. NM; Abad. GG: Notholt. J: Warneke. T; Hannigan. JW: Warneke. C: de
Gouw. JA; Dunlea. EJ; De Maziere. M; Griffith. DWT; Bernath. P; Jimenez. JL;
Wennberg. PO. (2011). Importance of secondary sources in the atmospheric budgets of
formic and acetic acids. Atmos Chem Phys 11: 1989-2013. http://dx.doi.org/ 10.5194/acp-
11-1989-2011.
Pavlik. M; Pavlikova. D; Zemanova. V: Hnilicka. F; Urbanova. V: Szakova. J. (2012). Trace
elements present in airborne particulate matter-Stressors of plant metabolism. Ecotoxicol
Environ Saf 79: 101-107. http://dx.doi.Org/10.1016/i.ecoenv.2011.12.009.
Payne. R; Gaud. V: Charm an. DJ. (2010). The impact of simulated sulfate deposition on
peatland testate amoebae. Microb Ecol 59: 76-83. http://dx.doi.org/10.1007/sQ0248-009-
9552-6.
Payne. RJ; Thompson. AM: Standen. V: Field. CD: Caporn. SJM. (2012). Impact of simulated
nitrogen pollution on heathland microfauna, mesofauna and plants. European Journal of
Soil Biology 49: 73-79. http://dx.doi.Org/10.1016/i.eisobi.2011.08.003.
Peel JL: Haeuber. R; Garcia. V: Russell AG: Neas. L. (2013). Impact of nitrogen and climate
change interactions on ambient air pollution and human health. Biogeochemistry 114:
121-134. http://dx.doi.org/10.1007/slQ533-012-9782-4.
Peierls. BL; Caraco. NF; Pace. ML: Cole. JJ. (1991). Human influence on river nitrogen. Nature
350: 386-387. http://dx.doi.org/10.1038/350386b0.
Peierls. BL: Paerl HW. (1997). Bioavailability of atmospheric organic nitrogen deposition to
coastal phytoplankton. Limnology 42: 1819-1823.
http://dx.doi.Org/10.4319/lo.1997.42.8.1819.
Pena. MA: Katsev. S: Oguz. T; Gilbert. D. (2010). Modeling dissolved oxygen dynamics and
hypoxia. Biogeosciences 7: 933-957. http://dx.doi.org/10.5194/bg-7-933-2010.
Peng. X: Jt O: Angell JH: Kearns. PJ: Yang. HJ: Bowen. JL: Ward. BB. (2016). Long-term
fertilization alters the relative importance of nitrate reduction pathways in salt marsh
sediments. Jour Geo Res: Biog 121: 2082-2095.
http://dx.doi.org/10.1002/2016JG0Q3484.
June 2018
R-169
DRAFT: Do Not Cite or Quote

-------
Peng. Y; Chen. G; Chen. G; Li. S; Peng. T; Qiu, X; Luo, J: Yang. S; Ftu, T; Ftu, H; Xu, Z; Liu.
L; Tang. Y; Tu. L. (2017). Soil biochemical responses to nitrogen addition in a secondary
evergreen broad-leaved forest ecosystem. Sci Rep 7: 2783.
http://dx.doi.org/10.1038/s41598-017-03Q44-w.
Pennanen. T; A. F; Fritze. H; Baath. E. (1996). Phospholipid fatty acid composition and heavy
metal tolerance of soil microbial communities along two heavy metal-polluted gradients
in coniferous forests. Appl Environ Microbiol 62: 420428.
Pennings. SC: Clark. CM: Cleland. EE: Collins. SL; Gough. L; Gross. KL; Milchunas. DG:
Suding. KN. (2005). Do individual plant species show predictable responses to nitrogen
addition across multiple experiments? Oikos 110: 547-555.
http://dx.doi.org/10.1111/j .0030-1299.2005.13792.x.
Perakis. SS: Sinkhorn. ER. (2011). Biogeochemistry of a temperate forest nitrogen gradient.
Ecology 92: 1481-1491. http://dx.doi.Org/10.1890/10-1642.l.
Perakis. SS: Sinkhorn. ER: Catricala. CE; Bullen. TP: Fitzpatrick. JA; Hynicka. JD; Cromack.
K. Jr. (2013). Forest calcium depletion and biotic retention along a soil nitrogen gradient.
Ecol Appl 23: 1947-1961. http://dx.doi.Org/10.1890/12-2204.l.
Perakis, SS: Sinkhorn, ER; Compton, JE. (2011). 815N constraints on long-term nitrogen
balances in temperate forests. Oecologia 167: 793-807. http://dx.doi.org/10.1007/sQ0442-
011-2016-v.
Percent, SF; Frischer, ME: Vescio, PA: Duffy, EB; Milano, V: Mclellan, M; Stevens, BM;
Boylen, CW; Nierzwicki-Bauer, SA. (2008). Bacterial community structure of acid-
impacted lakes: What controls diversity? Appl Environ Microbiol 74: 1856-1868.
http://dx.doi.org/10.1128/AEM.01719-07.
Peteriohn, WT: Adams, MB: Gilliam, FS. (1996). Symptoms of nitrogen saturation in two
central Appalachian hardwood forest ecosystems. Biogeochemistry 35: 507-522.
http://dx.doi. org/10.1007/BF02183038.
Petersen, JK: Hasler, B: Timmermann, K: Nielsen, P; Torring, DB: Larsen, MM: Holmer, M.
(2014). Mussels as a tool for mitigation of nutrients in the marine environment. Mar
PollutBull 82: 137-143. http://dx.doi.Org/10.1016/i.marpolbul.2014.03.006.
Petersen, JK: Saurel, C: Nielsen, P; Timmermann, K. (2016). The use of shellfish for
eutrophication control. Aquaculture International 24: 857-878.
http://dx.doi.org/10.1007/slQ499-015-9953-0.
Peterson, RH: Martin-Robichaud, DJ. (1986). Growth and major inorganic cation budgets of
Atlantic salmon alevins at three ambient acidities. Trans Am Fish Soc 115: 220-226.
http://dx.doi.org/10.1577/1548-8659(1986")! 15<220:GAMICB>2.0.CQ:2.
Peterson, RH: Van Eeckhaute, L. (1992). Distributions of Ephemeroptera, Plecoptera, and
Trichoptera of three maritime catchments differing in pH. Freshw Biol 27: 65-78.
http://dx.doi.Org/10.llll/i.1365-2427.1992.tb00523.x.
Petraitis, PS: Latham, RE. (1999). The importance of scale in testing the origins of alternative
community states. Ecology 80: 429-442. http://dx.doi.org/10.1890/0Q12-
9658n999X)80r0429:TIOSIT12.0.CO:2.
June 2018
R-170
DRAFT: Do Not Cite or Quote

-------
Petroff. A; Mailliat. A; Amielh. M; Anselmet F. (2008). Aerosol dry deposition on vegetative
canopies. Part I: Review of present knowledge. Atmos Environ 42: 3625-3653.
http://dx.doi.ore/10.1016/i.atmosenv.2007.09.043.
Pfister. CA; Esbaugh, AJ; Frieder. CA: Baumann. H: Bockmon. EE: White. MM: Carter. BR:
Benwav. HM; Blanchette. CA: Carrington, E: Mcclintock. JB: Mccorkle. DC: Mcgillis.
WR: Moonev. TA: Ziveri. P. (2014). Detecting theunexpected: A research framework for
ocean acidification [Review], Environ Sci Technol 48: 9982-9994.
http://dx.doi.org/10.1021/es5Q1936p.
Phelan. J: Belvazid. S: Jones. P: Caika. J: Buckley. J: Clark. C. (2016). Assessing the effects of
climate change and air pollution on soil properties and plant diversity in sugar maple-
beech-yellow birch hardwood forests in the northeastern United States: Model
simulations from 1900 to 2100. Water Air Soil Pollut 227: 84.
http://dx.doi. org/10.1007/s 11270-016-2762-x.
Phelan. J: Belvazid. S: Kurz. D: Guthrie. S: Caika. J: Sverdrup. Ft: Waite. R. (2014). Estimation
of soil base cation weathering rates with the PROFILE model to determine critical loads
of acidity for forested ecosystems in Pennsylvania, USA: Pilot application of a potential
national methodology. Water Air Soil Pollut 225: 2109-2128.
http://dx.doi.org/10.1007/sll270-014-21Q9-4.
Phoenix, GK; Emmett, BA; Britton, AJ: Caporn, SJM; Pise, NB; Helliwell, R; Jones, L; Leake,
JR: Leith, ID: Sheppard, LJ: Sowerbv, A: Pilkington, MG: Rowe, EC: Ashmorek, MR:
Power, SA. (2012). Impacts of atmospheric nitrogen deposition: Responses of multiple
plant and soil parameters across contrasting ecosystems in long-term field experiments.
Global Change Biol 18: 1197-1215. http://dx.doi.org/10.1111/i. 1365-2486.2011,02590.x.
Phoenix, GK: Hicks, WK: Cinderbv, S: Kuvlenstierna, JCI; Stock, WD: Dentener, FJ: Giller,
KE: Austin, AT: Lefrov, RDB: Gimeno, BS: Ashmore, MR: Ineson, P. (2006).
Atmospheric nitrogen deposition in world biodiversity hotspots: The need for a greater
global perspective in assessing N deposition impacts. Global Change Biol 12: 470-476.
http://dx.doi.org/10.1111/i. 1365-2486.2006.01104.x.
Pickett, STA: Cadenasso, ML: Grove, JM: Boone, CG: Groffman, PM: Irwin, E: Kaushal, SS:
Marshall, V: McGrath, BP: Nilon, CH: Pouvat, RV: Szlavecz, K: Troy, A: Warren, P.
(2011). Urban ecological systems: Scientific foundations and a decade of progress. J
Environ Manage 92: 331-362. http://dx.doi.Org/10.1016/i.ienvman.2010.08.022.
Piehler, MF; Smyth, AR. (2011). Habitat-specific distinctions in estuarine denitrification affect
both ecosystem function and services. Ecosphere 2. http://dx.doi.org/10.189Q/ES10-
00082.1.
Piessens, K: Honnav, O: Nackaerts, K: Hermy, M. (2004). Plant species richness and
composition of heathland relics in north-western Belgium: Evidence for a rescue-effect?
Journal of Biogeography 31: 1683-1692. http://dx.doi.Org/10.l 111/i. 1365-
2699.2004.01056.x.
Pieterse, G: Bleeker, A: Vermeulen, AT: Wu, Y; Erisman, JW. (2007). High resolution
modelling of atmosphere-canopy exchange of acidifying and eutrophying components
and carbon dioxide for European forests. Tellus B Chem Phys Meteorol 59: 412-424.
http://dx.doi.org/10.1111/i. 1600-0889.2007.00266.X.
June 2018
R-171
DRAFT: Do Not Cite or Quote

-------
Pimentel, D; Wilson. C; McCullum, C; Huang. R; Dwen, P; Flack. J; Iran. 0; Saltman, T; Cliff.
R. (1997). Economic and environmental benefits of biodiversity. Bioscience 47: 747-757.
http://dx.doi.org/10.2307/1313Q97.
Pinder. RW: Bettez. ND; Bonan. GB; Greaver. TL; Wieder. WR; Schlesinger. WH; Davidson.
EA. (2013). Impacts of human alteration of the nitrogen cycle in the US on radiative
forcing. Biogeochemistry 114: 25-40. http://dx.doi.org/10.1007/slQ533-012-9787-z.
Pinder. RW: Davidson. EA: Goodale. CL; Greaver. TL: Herrick. JD; Liu. L. (2012). Climate
change impacts of US reactive nitrogen. Proc Natl Acad Sci USA 109: 7671-7675.
http://dx.doi.org/10.1073/pnas. 1114243109.
Pinder. RW: Gilliland. AB; Dennis. RL. (2008). Environmental impact of atmospheric NH3
emissions under present and future conditions in the eastern United States. Geophys Res
Lett 35. http://dx.doi.org/10.1029/2008GLQ33732.
Pinder, RW: Pekney, NJ; Davidson, CI; Adams, PJ. (2004a). A process-based model of ammonia
emissions from dairy cows: improved temporal and spatial resolution. Atmos Environ 38:
1357-1365. http://dx.doi.Org/10.1016/i.atmosenv.2003.l 1.024.
Pinder, RW; Strader, R; Davidson, CI; Adams, PJ. (2004b). A temporally and spatially resolved
ammonia emission inventory for dairy cows in the United States. Atmos Environ 38:
3747-3756. http://dx.doi.Org/10.1016/i.atmosenv.2004.04.008.
Pinder, RW; Walker, JT; Bash, JO; Cadv-Pereira, KE; Henze, DK; Luo, M; Osterman, GB;
Shephard, MW. (2011). Quantifying spatial and seasonal variability in atmospheric
ammonia with in situ and space-based observations. Geophys Res Lett 38: L04802.
http://dx.doi.org/10.1029/2010GLQ46146.
Pinsonneault, AJ; Moore, T, imR; Roulet, NT. (2016). Effects of long-term fertilization on peat
stoichiometry and associated microbial enzyme activity in an ombrotrophic bog.
Biogeochemistry 129: 149-164. http://dx.doi.org/10.1007/slQ533-016-0224-6.
Pinto, JP; Dibb, J; Lee, BH; Rappengltick, B; Wood, EC; Levy, M; Zhang, RY; Lefer, B; Ren,
XR; Stutz, J; Tsai, C; Ackermann, L; Golovko, J; Herndon, SC; Oakes, M; Meng, QY;
Munger, JW; Zahniser, M; Zheng, J. (2014). Intercomparison of field measurements of
nitrous acid (HONO) during the SHARP campaign. J Geophys Res Atmos 119: 5583-
5601. http://dx.doi.org/10.1002/2013JD02Q287.
Pirrone, N; Cinnirella, S; Feng, X; Finkelman, RB; Friedli, HR; Leaner, J; Mason, R; Mukheriee,
AB; Stracher, GB; Streets, DG; Telmer, K. (2010). Global mercury emissions to the
atmosphere from anthropogenic and natural sources. Atmos Chem Phys 10: 5951-5964.
http://dx.doi.org/10.5194/acp-10-5951-2010.
Pitel, NE; Yanai, RD. (2014). Abiotic and biotic factors influencing sugar maple health: Soils,
topography, climate, and defoliation. Soil Sci Soc Am J 78: 2061-2070.
http://dx.doi.org/10.2136/sssai2014.06.024Q.
Pither, J; Aarssen, LW. (2006). How prevalent are pH-specialist diatoms? A reply to Telford et
al. (2006). Ecol Lett 9: E6-E12. http://dx.doi.Org/10.l 111/i. 1461-0248.2006.00906.X.
June 2018
R-172
DRAFT: Do Not Cite or Quote

-------
Pivovaroff. AL; Santiago. LS; Vourlitis. GL; Grantz. DA; Allen. MF. (2016). Plant hydraulic
responses to long-term dry season nitrogen deposition alter drought tolerance in a
Mediterranean-type ecosystem. Oecologia 181: 721-731.
http://dx.doi. org/10.1007/s00442-016-3609-2.
Placet. M; Battve. RE: Barnard. WR; Gillette. DA: Johnson. TC. (1990). Emissions involved in
acidic deposition processes. In Acid Deposition: State of Science and Technology: Report
1. Washington, DC: The U.S. National Acid Precipitation Assessment Program.
Piatt. U: Stutz. J. (2008). Physics of earth and space environments: Differential optical
absorption spectroscopy: Principles and applications. In Physics of Earth and Space
Environments (1 ed.). Berlin, Germany: Springer-Verlag. http://dx.doi.org/10.10Q7/978-
3-540-75776-4.
Pleim. J: Ran. L. (2011). Surface flux modeling for air quality applications. Atmosphere (Basel)
2: 271-302. http://dx.doi.org/10.3390/atmos2030271.
Plummer. ML: Harvey. CJ: Anderson. LE; Guerry. AD: Ruckelshaus. MH. (2013). The role of
eelgrass in marine community interactions and ecosystem services: Results from
ecosystem-scale food web models. Ecosystems 16: 237-251.
http://dx.doi.org/10.1007/sl0021-012-9609-Q.
Plummer. P; Tobias. C: Cadv. D. (2015). Nitrogen reduction pathways in estuarine sediments:
Influences of organic carbon and sulfide. Jour Geo Res: Biog 120: 1958-1972.
http://dx.doi.org/10.1002/2015JG003Q57.
Plutchak, R; Major, K; Cebrian, J: Foster, CD: Miller, MEC; Anton, A; Sheehan, KL; Heck, KL,
Jr; Powers, SP. (2010). Impacts of oyster reef restoration on primary productivity and
nutrient dynamics in tidal creeks of the north central Gulf of Mexico. Estuaries Coasts
33: 1355-1364. http://dx.doi.org/10.1007/sl2237-01Q-9327-9.
Poland, TM; McCullough, DG. (2006). Emerald ash borer: Invasion of the urban forest and the
threat to North America's ash resource. J Forest 104: 118-124.
Pollack, JB; Yoskowitz, D; Kim, HC: Montagna, PA. (2013). Role and value of nitrogen
regulation provided by oysters (Crassostrea virginica) in the Mission-Aransas Estuary,
Texas, USA. PLoS ONE 8: e65314. http://dx.doi.org/10.1371/iournal.pone.0065314.
Pollman, CD. (2014). Mercury cycling in aquatic ecosystems and trophic state-related variables -
Implications from structural equation modeling. Sci Total Environ 499: 62-73.
http://dx.doi.Org/10.1016/i.scitotenv.2014.08.036.
Polo-Cavia, N: Burraco, P; Gomez-Mestre, I. (2016). Low levels of chemical anthropogenic
pollution may threaten amphibians by impairing predator recognition. Aquat Toxicol
172: 30-35. http://dx.doi.Org/10.1016/i.aquatox.2015.12.019.
Poor, ND; Cross, LM; Dennis, RL. (2013a). Lessons learned from the Bay Region Atmospheric
Chemistry Experiment (BRACE) and implications for nitrogen management of Tampa
Bay. Atmos Environ 70: 75-83. http://dx.doi.Org/10.1016/i.atmosenv.2012.12.030.
June 2018
R-173
DRAFT: Do Not Cite or Quote

-------
Poor. ND; Pribble. JR; Schwede. DB. (2013b). Application of watershed deposition tool to
estimate from CMAQ simulations the atmospheric deposition of nitrogen to Tampa Bay
and its watershed. J Air Waste Manag Assoc 63: 100-114.
http://dx.doi.org/10.1080/10962247.2012.7391Q9.
Popek. R; Gawronska. H; Gawronski. SW. (2015). The Level of Particulate Matter on Foliage
Depends on the Distance from the Source of Emission. Int J Phytoremediation 17: 1262-
1268. http://dx.doi.org/10.1080/15226514.2014.989312.
Popek. R; Gawronska. H; Wrochna. M; Gawronski. SW: Saebo. A. (2013). Particulate matter on
foliage of 13 woody species: Deposition on surfaces and phytostabilisation in waxes - a
3-year study. Int J Phytoremediation 15: 245-256.
http://dx.doi.org/10.1080/15226514.2012.694498.
Popek. R; Lukowski. A: Karolewski. P. (2017). Particulate matter accumulation - further
differences between native Prunus padus and non-native P.serotina. Dendrobiology 78:
85-95. http://dx.doi.org/10.12657/denbio.078.009.
Porcal. P; Koprivniak. JF; Molot. LA: Dillon. PJ. (2009). Humic substances - part 7: the
biogeochemistry of dissolved organic carbon and its interactions with climate change
[Review], Environ Sci Pollut Res Int 16: 714-726. http://dx.doi.org/10.1007/sl 1356-009-
0176-7.
Porter. E; Blett. T; Potter. DU; Huber. C. (2005). Protecting resources on federal lands:
Implications of critical loads for atmospheric deposition of nitrogen and sulfur.
Bioscience 55: 603-612. http://dx.doi.org/10.1641/00Q6-
3 568(2005^)055 r0603:PROFLn2.0.CQ:2.
Porter. E; Johnson. S. (2007). Translating science into policy: Using ecosystem thresholds to
protect resources in Rocky Mountain National Park. Environ Pollut 149: 268-280.
http://dx.doi.Org/10.1016/i.envpol.2007.06.060.
Porter. EM: Bowman. WD: Clark. CM: Compton. JE; Pardo. LH; Soong. JL. (2013). Interactive
effects of anthropogenic nitrogen enrichment and climate change on terrestrial and
aquatic biodiversity. Biogeochemistry 114: 93-120. http://dx.doi.org/10.1007/slQ533-
012-9803-3.
Posch. M; Aherne. J: Hettelingh. JP. (2011). Nitrogen critical loads using biodiversity-related
critical limits. Environ Pollut 159: 2223-2227.
http://dx.doi.Org/10.1016/i.envpol.2010.l 1.001.
Posch, M; Reinds, GJ. (2009). A very simple dynamic soil acidification model for scenario
analyses and target load calculations. Environ Modell Softw 24: 329-340.
http://dx.doi.Org/10.1016/i.envsoft.2008.09.007.
Poulin. BA; Aiken. GR; Nagy. KL; Monceau. A: Krabbenhoft. DP: Ryan. JN. (2016). Mercury
transformation and release differs with depth and time in a contaminated riparian
soil during simulated flooding. Geochim Cosmo Act 176: 118-138.
June 2018
R-174
DRAFT: Do Not Cite or Quote

-------
Pourmokhtarian. A; Driscoll. CT; Campbell. JL; Havhoe. K. (2012). Modeling potential
hydrochemical responses to climate change and increasing C02 at the Hubbard Brook
Experimental Forest using a dynamic biogeochemical model (PnET-BGC). Water Resour
Res 48. http://dx.doi.org/10.1029/2Ql 1WR011228.
Povak. NA; Hessburg. PF; Mcdonnell. TC; Reynolds. KM; Sullivan. TJ; Salter. RB; Cosby. BJ.
(2014). Machine learning and linear regression models to predict catchment- level base
cation weathering rates across the southern Appalachian Mountain region, USA. Water
Resour Res 50: 2798-2814. http://dx.doi.org/10.1002/2013WR0142Q3.
Povak. NA: Hessburg. PF: Reynolds. KM: Sullivan. TJ: Mcdonnell. TC: Salter. RB. (2013).
Machine learning and hurdle models for improving regional predictions of stream water
acid neutralizing capacity. Water Resour Res 49: 3531-3546.
http://dx.doi.org/10.1002/wrcr.203Q8.
Powell. AS; Jackson. L; Ardon. M. (2016). Disentangling the effects of drought, salinity, and
sulfate on baldcypress growth in a coastal plain restored wetland. Restor Ecol 24: 548-
557. http://dx.doi.org/10. Ill 1/rec. 12349.
Poxleitner. M; Trommer. G; Lorenz. P; Stibor. H. (2016). The effect of increased nitrogen load
on phytoplankton in a phosphorus-limited lake. FreshwBiol 61: 1966-1980.
http://dx.doi.org/10. Ill 1/fwb. 12829.
Prasad. MNV; DeOliveira. FHM. (2003). Metal hyperaccumulation in plants: Biodiversity
prospecting for phytoremediation technology. Electron J Biotechnol 6: 110-146.
Preece, EP; Hardy, FJ; Moore, BC; Bryan, M. (2017). A review of microcystin detections in
Estuarine and Marine waters: Environmental implications and human health risk.
Harmful Algae 61: 31-45. http://dx.doi.Org/10.1016/i.hal.2016.l 1.006.
Pregitzer, KS; Burton, AJ; Zak, PR; Talhelm, AF. (2008). Simulated chronic nitrogen deposition
increases carbon storage in Northern Temperate forests. Global Change Biol 14: 142-153.
http://dx.doi.org/10.111 l/i.l365-2486.2007.01465.x.
PREP (Piscataqua Region Estuaries Partnership). (2013). State of Our Estuaries 2013.
Preston, SD; Alexander, RB; Schwarz, GE; Crawford, CG. (2011). FACTORS AFFECTING
STREAM NUTRIENT LOADS: A SYNTHESIS OF REGIONAL SPARROW MODEL
RESULTS FOR THE CONTINENTAL UNITED STATES. J Am Water Resour Assoc
47: 891-915. http://dx.doi.Org/10.l 111/i. 1752-1688.2011.00577.x.
Pritchard, SG; Taylor, BN; Cooper, ER; Beidler, KV; Strand, AE; Mccormack, ML; Zhang, S.
(2014). Long-term dynamics of mycorrhizal root tips in a loblolly pine forest grown with
free-air C02 enrichment and soil N fertilization for 6 years. Global Change Biol 20:
1313-1326. http://dx.doi.org/10.1111/gcb. 12409.
Pritz. CMF; Schrlau. JE; Simonich. SLM; Blett. TF. (2014). CONTAMINANTS OF
EMERGING CONCERN IN FISH FROM WESTERN US AND ALASKAN
NATIONAL PARKS - SPATIAL DISTRIBUTION AND HEALTH THRESHOLDS. J
Am Water Resour Assoc 50: 309-323. http://dx.doi.0rg/lO.l 111/iawr. 12168.
June 2018
R-175
DRAFT: Do Not Cite or Quote

-------
Providoli. I; Bugmann. H; Siegwolf. R: Buchmann. N; Schleppi. P. (2005). Flow of deposited
inorganic N in two Gleysol-dominated mountain catchments traced with 15N03- and
15NH4+. Biogeochemistry 76: 453-475. http://dx.doi.org/10.1007/slQ533-005-8124-l.
Przybvsz. A: Popek. R: Gawronska. H; Grab. K; Loskot K: Wrochna. M: Gawronski. SW.
(2014). Efficiency of photosynthetic apparatus of plants grown in sites differing in level
of particulate matter. Acta Scientiarum Polonorum Hortorum Cultus 13: 17-30.
Puchalski. MA: Rogers. CM: Baumgardner. R: Mishoe. KP: Price. G: Smith. MJ; Watkins. N:
Lehmann. CM. (2015). A statistical comparison of active and passive ammonia
measurements collected at Clean Air Status and Trends Network (CASTNET) sites.
Environ Sci Process Impacts 17: 358-369. http://dx.doi.org/10.1039/c4em00531g.
Puchalski. MA: Sather. ME: Walker. JT; Lehmann. CMB; Gay. DA: Mathew. J: Robarge. WP.
(2011). Passive ammonia monitoring in the United States: Comparing three different
sampling devices. J Environ Monit 13: 3156-3167.
http://dx.doi.org/10.1039/cleml0553a.
Pulido. C: Riera. JL: Ballesteros. E: Chappuis. E: Gacia. E. (2015). Predicting aquatic
macrophyte occurrence in soft-water oligotrophic lakes (Pyrenees mountain range).
Journal of Limnology 74: 143-154. http://dx.doi.org/10.4081/ilimnol.2014.965.
Qiao. Z; Murray. F. (1998). Improvement of the 15N dilution method for estimation of
absorption of NOx by plants supplied with 15N-labelled fertilizer. New Phytol 138: 13-
18.
Qu. J: Ren. G: Chen. B; Fan. J: E. Y. (2011). Effects of lead and zinc mining contamination on
bacterial community diversity and enzyme activities of vicinal cropland. Environ Monit
Assess 182: 597-606. http://dx.doi.org/10.1007/slQ661-011-1900-6.
Quesnel. P. -O: Cote. B. (2009). Prevalence of phosphorus, potassium, and calcium limitations in
white spruce across Canada. J Plant Nutr 32: 1290-1305.
http://dx.doi.org/10.1080/019041609030060Q2.
R'Honi. Y; Clarisse. L; Clerbaux. C: Hurtmans. D; Duflot. V: Turquetv. S: Ngadi. Y; Coheur.
PF. (2013). Exceptional emissions of NH3 and HCOOH in the 2010 Russian wildfires.
Atmos Chem Phys 13: 4171-4181. http://dx.doi.org/10.5194/acp-13-4171-2013.
Rabalais. NN. (2002). Nitrogen in aquatic ecosystems. Ambio 31: 102-112.
http://dx.doi.org/10.1639/0044-7447(2002X)3ir0102:NIAE12.0.CO;2.
Rabalais. NN: Diaz. RJ; Levin. LA: Turner. RE: Gilbert. D; Zhang. J. (2010). Dynamics and
distribution of natural and human-caused hypoxia. Biogeosciences 7: 585-619.
http://dx.doi.org/10.5194/bg-7-585-2010.
Rabalais. NN: Turner. RE: Diaz. RJ: Justic. D. (2009). Global change and eutrophication of
coastal waters. ICES J Mar Sci 66: 1528-1537. http://dx.doi.org/10.1093/icesims/fsp047.
Raddum. GG: Brettum. P: Matzow. D: Nilssen. JP: Skov. A: Svealv. T: Wright. RF. (1986).
Liming the acid lake Hovvatn, Norway: A whole-ecosystem study. Water Air Soil Pollut
31: 721-763. http://dx.doi.org/10.1007/BF00284222.
June 2018
R-176
DRAFT: Do Not Cite or Quote

-------
Rago, PJ: Wiener. JG. (1986). Does pH affect fish species richness when lake area is considered?
Trans Am Fish Soc 115: 438-447. http://dx.doi.org/10.1577/1548-
8659(1986)115<438:DPAFSR>2.0.CO:2.
Rahman. MS: Thomas. P. (2015). Molecular characterization and hypoxia-induced upregulation
of neuronal nitric oxide synthase in Atlantic croaker: Reversal by antioxidant and
estrogen treatments. Comp Biochem Physiol A Mol Integr Physiol 185: 91-106.
http://dx.doi.Org/10.1016/i.cbpa.2015.03.013.
Rai. A: Kulshreshtha. K; Srivastava. PK; Mohantv. CS. (2010). Leaf surface structure alterations
due to particulate pollution in some common plants. Environmentalist 30: 18-23.
http://dx.doi.org/10.1007/slQ669-009-9238-0.
Rai a. R; Navak. AK; Rao. KS: Puree. C: Shahid. M; Panda. BB; Kumar. A: Tripathi. R;
Bhattacharyya. P; Baig. MJ; Lai. B; Mohantv. S: Gautam. P. (2014). Effect of fly ash
deposition on photosynthesis, growth and yield of rice. Bull Environ Contam Toxicol 93:
106-112. http://dx.doi.org/10.1007/sQ0128-014-1282-x.
Ramial. PS: Rudd. JWM; Furutam. A: Xun. L. (1985). The effect of pH on methyl mercury
production and decomposition in lake sediments. Can J Fish Aquat Sci 42: 685-692.
http://dx.doi.Org/10.l 139/f85-088.
Ramirez. KS: Craine. JM; Fierer. N. (2010a). Nitrogen fertilization inhibits soil microbial
respiration regardless of the form of nitrogen applied. Soil Biol Biochem 42: 2336-2338.
http://dx.doi.Org/10.1016/i.soilbio.2010.08.032.
Ramirez, KS: Lauber, CL; Knight, R; Bradford, MA; Fierer, N. (2010b). Consistent effects of
nitrogen fertilization on soil bacterial communities in contrasting systems. Ecology 91:
3463-3470. http://dx.doi.Org/10.1890/10-0426.l.
Ramial, PS: Rudd, JWM: Hecky, RE. (1986). Methods for measuring specific rates of mercury
methylation and degradation and their use in determining factors controlling net rates of
mercury methylation. Appl Environ Microbiol 51: 110-114.
Rao, LE; Allen, EB. (2010). Combined effects of precipitation and nitrogen deposition on native
and invasive winter annual production in California deserts. Oecologia 162: 1035-1046.
http://dx.doi. org/10.1007/s00442-009-1516-5.
Rao, LE: Allen, EB: Meixner, T. (2010). Risk-based determination of critical nitrogen deposition
loads for fire spread in southern California deserts. Ecol Appl 20: 1320-1335.
http://dx.doi.Org/10.1890/09-0398.l.
Rao, LE: Matchett, JR: Brooks, ML: Johnson, RF: Minnich, RA: Allen, EB. (2015).
Relationships between annual plant productivity, nitrogen deposition and fire size in low-
elevation California desert scrub. International Journal of Wildland Fire 24: 48-58.
http://dx.doi.org/10.1071/WF13152.
Rao, LE: Parker, PR: Bytnerowicz, A: Allen, EB. (2009). Nitrogen mineralization across an
atmospheric nitrogen deposition gradient in Southern California deserts. J Arid Environ
73: 920-930. http://dx.doi.Org/10.1016/i.iaridenv.2009.04.007.
June 2018
R-177
DRAFT: Do Not Cite or Quote

-------
Rao. LE; Steers. RJ; Allen. EB. (2011). Effects of natural and anthropogenic gradients on native
and exotic winter annuals in a southern California Desert. Plant Ecol 212: 1079-1089.
http://dx.doi.org/10.1007/sll258-01Q-9888-5.
Rao. P; Hutyra. LR; Raciti. SM; Templer. PH. (2014). Atmospheric nitrogen inputs and losses
along an urbanization gradient from Boston to Harvard Forest, MA. Biogeochemistry
121: 229-245. http://dx.doi.org/10.1007/slQ533-013-9861-l.
Rasanen. JV; Holopainen. T; Joutsensaari. J: Pasanen. P; Kivimaenpaa. M. (2014). Particle
capture efficiency of different-aged needles of Norway spruce under moderate and severe
drought. Can J For Res 44: 831-835. http://dx.doi.Org/10.l 139/cifr-2014-0068.
Rask. M; Vuorenmaa. J: Nvberg. K; Tammi. J: Mannio. J: Olin. M; Kortelainen. P; Raitaniemi.
J: Vesala. S. (2014). Recovery of acidified lakes in Finland and subsequent responses of
perch and roach populations. Boreal Environmental Research 19: 222-234.
Rea, AW; Davis, C; Evans, DA; Heninger, BT; Van Houtven, G. (2012). Using ecosystem
services to inform decisions on U.S. air quality standards [Review], Environ Sci Technol
46: 6481-6488. http://dx.doi.org/10.1021/es3003876.
Read, DJ; Leake, JR; Perez-Moreno, J. (2004). Mycorrhizal fungi as drivers of ecosystem
processes in heathland and boreal forest biomes. Can J Bot 82: 1243-1263.
http ://dx. doi .org/10.113 9/B04-123.
Reader, JP; Dalziel, TRK; Morris, R. (1988). Growth, mineral uptake and skeletal calcium
deposition in brown trout, Salmo trutta L, yolk-sac fry exposed to aluminum and
manganese in soft acid water. J Fish Biol 32: 607-624. http://dx.doi.org/10.1111/j .1095-
8649.1988.tb05399.x.
Reclamation, USBO; MPR; Service, M. (1996). Digitial representation of Kiichler vegetation
potential map for California: Natural Vegetation of California. U.S. Bureau Of
Reclamation; MPR; MPGIS Service.
Redbo-Torstensson, P. (1994). The demographic consequences of nitrogen fertilization of a
population of sundew, Drosera rotundifolia. Acta Bot Neerl 43: 175-188.
http://dx.doi.Org/10.llll/i.1438-8677.1994.tb00743.x.
Reddv, KR; Patrick, WH. (1984). Nitrogen transformations and loss in flooded soils and
sediments. Crit Rev Environ Sci Tech 13: 273-309.
Reed, ML; Pinckney, JL; Keppler, CJ; Brock, LM; Hogan, SB; Greenfield, DI. (2016). The
influence of nitrogen and phosphorus on phytoplankton growth and assemblage
composition in four coastal, southeastern USA systems. Estuar Coast Shelf Sci 177: 71-
82. http://dx.doi.Org/10.1016/i.ecss.2016.05.002.
Regoli, F; Gorbi, S; Fattorini, D; Tedesco, S; Notti, A; Machella, N; Bocchetti, R; Benedetti, M;
Piva, F. (2006). Use of the land snail Helix aspersa sentinel organism for monitoring
ecotoxicologic effects of urban pollution: An integrated approach. Environ Health
Perspect 114: 63-69. http://dx.doi.org/10.1289/ehp.8397.
June 2018
R-178
DRAFT: Do Not Cite or Quote

-------
Regus, JU; Wendlandt CE; Bantav. RM; Gano-Cohen. KA; Gleason. NJ; Hollo well. AC;
O'Neill MR; Shahin. KK; Sachs. JL. (2017). Nitrogen deposition decreases the benefits
of symbiosis in a native legume. Plant Soil 414: 159-170.
http://dx.doi.org/10.1007/sl 1104-016-3114-8.
Reich. PB. (2009). Elevated C02 reduces losses of plant diversity caused by nitrogen deposition.
Science 326: 1399-1402. http://dx.doi.Org/10.l 126/science.l 178820.
Reich. PB; Buschena. C; Tioelker. MG; Wrage. K; Knops. J; Tilman. D; Machado. JL. (2003).
Variation in growth rate and ecophysiology among 34 grassland and savanna species
under contrasting N supply: A test of functional group differences. New Phytol 157: 617-
631. http://dx.doi.Org/10.1046/i.1469-8137.2003.00703.x.
Reich. PB; Hobbie. SE; Lee. TP. (2014). Plant growth enhancement by elevated C02 eliminated
by joint water and nitrogen limitation. Nat Geosci 7: 920-924.
http://dx.doi.org/10.1038/NGEQ2284.
Reich. PB; Knops. J; Tilman. D; Craine. J; Ellsworth. D; Tioelker. M; Lee. T; Wedin. D; Naeem.
S; Bahauddin. D; Hendrev. G; Jose. S; Wrage. K; Goth. J; Bengston. W. (2001). Plant
diversity enhances ecosystem responses to elevated C02 and nitrogen deposition. Nature
410: 809-812. http://dx.doi.org/10.1038/35071062.
Reich. PB; Tilman. D; Isbell. F; Mueller. K; Hobbie. SE; Flynn. DF; Eisenhauer. N. (2012).
Impacts of biodiversity loss escalate through time as redundancy fades. Science 336: 589-
592. http://dx.doi.Org/10.l 126/science. 1217909.
Reich, PB; Tioelker, MG; Machado, JL; Oleksyn, J. (2006). Universal scaling of respiratory
metabolism, size and nitrogen in plants. Nature 439: 457-461.
http://dx.doi.org/10.1038/nature04282.
Reich, PB; Tioelker, MG; Pregitzer, KS; Wright, I, anJ; Oleksyn, J; Machado, JL. (2008).
Scaling of respiration to nitrogen in leaves, stems and roots of higher land plants. Ecol
Lett 11: 793-801. http://dx.doi.org/10.1111/i. 1461-0248.2008.01185.x.
Reid, CR; Watmough, SA. (2016). Spatial patterns, trends, and the potential long-term impacts
of tree harvesting on lake calcium levels in the Muskoka River Watershed, Ontario,
Canada. Can J Fish Aquat Sci 73: 382-393. http://dx.doi.org/ 10.1139/cifas-2015-0231.
Reinds, G; de Vries, W. (2010). Uncertainties in critical loads and target loads of sulphur and
nitrogen for European forests: Analysis and quantification. Sci Total Environ 408: 1960-
1970. http://dx.doi.Org/10.1016/i.scitotenv.2009.12.001.
Reinfelder, JR; Fisher, NS; Luoma, SN; Nichols, JW; Wang, WX. (1998). Trace element trophic
transfer in aquatic organisms: A critique of the kinetic model approach. Sci Total Environ
219: 117-135. http://dx.doi.org/10.1016/S0048-9697(98X)0225-3.
Reisinger, AJ; Tank, JL; Dee, MM. (2016). Regional and seasonal variation in nutrient limitation
of river biofilms. Freshw Sci 35: 474-489. http://dx.doi.org/10.1086/685829.
Reitsma, J; Murphy, DC; Archer, AF; York, RH. (2017). Nitrogen extraction potential of wild
and cultured bivalves harvested from nearshore waters of Cape Cod, USA. Mar Pollut
Bull 116: 175-181. http://dx.doi.Org/10.1016/i.marpolbul.2016.12.072.
June 2018
R-179
DRAFT: Do Not Cite or Quote

-------
Ren. H; Chen. YC: Wang. XT; Wong. GTF; Cohen. AL; DeCarlo. TM; Weigand. MA; Mii. HS;
Sigman. DM. (2017). 21st-century rise in anthropogenic nitrogen deposition on a remote
coral reef. Science 356: 749-752. http://dx.doi.Org/10.l 126/science.aal3869.
Reuss. JO. (1983). Implications of the calcium-aluminum exchange system for the effect of acid
precipitation on soils. J Environ Qual 12: 591-595.
http://dx.doi.org/10.2134/ieal983.00472425001200040Q32x.
Reuss. JO; Johnson. DW. (1985). Effect of soil processes on the acidification of water by acid
deposition. J Environ Qual 14: 26-31.
Reuss, JO; Johnson, DW. (1986). Acid deposition and the acidification of soils and waters. In
Acid Deposition and the Acidification of Soils and Waters. New York, NY: Springer-
Verlag New York, http://dx.doi.org/10.1007/978-l-4419-8536-l.
Revmond, CE; Lloyd, A; Kline, DI; Dove, SG; Pandolfi, JM. (2013). Decline in growth of
foraminifer Marginopora rossi under eutrophication and ocean acidification scenarios.
Global Change Biol 19: 291-302. http://dx.doi.org/10. Ill 1/gcb. 12035.
Reynolds, PL; Richardson, JP; Duffy, JE. (2014). Field experimental evidence that grazers
mediate transition between microalgal and seagrass dominance. Limnol Oceanogr 59:
1053-1064. http://dx.doi.Org/10.4319/lo.2014.59.3.1053.
Rheuban, JE; Williamson, S; Costa, JE; Glover, DM; Jakuba, RW; Mccorkle, DC; Neill, C;
Williams, T; Donev, SC. (2016). Spatial and temporal trends in summertime climate and
water quality indicators in the coastal embayments of Buzzards Bay, Massachusetts.
Biogeosciences 13: 253-265. http://dx.doi.org/10.5194/bg-13-253-2016.
Rhoades, CC; Mccutchan, JH, Jr; Cooper, LA; Clow, D; Detmer, TM; Briggs, JS; Stednick, JD;
Veblen, TT; Ertz, RM; Likens, GE; Lewis, WM, Jr. (2013). Biogeochemistry of beetle-
killed forests: Explaining a weak nitrate response. Proc Natl Acad Sci USA 110: 1756-
1760. http://dx.doi.org/10.1073/pnas.1221029110.
Rhodes, C; Bingham, A; Heard, AM; Hewitt, J; Lynch, J; Waite, R; Bell, MP. (2017). Diatoms
to human uses: linking nitrogen deposition, aquatic eutrophi cation, and ecosystem
services. Ecosphere 8. http://dx.doi.org/10.10Q2/ecs2.1858.
Rice, KC; Scanlon, TM; Lynch, JA; Cosby, BJ. (2014). Decreased atmospheric sulfur deposition
across the Southeastern U.S.: When will watersheds release stored sulfate? Environ Sci
Technol 48: 10071-10078. http://dx.doi.org/10.1021/es5Q1579s.
Rich, JJ; Dale, OR; Song, B; Ward, BB. (2008). Anaerobic ammonium oxidation (anammox) in
Chesapeake Bay sediments. Microb Ecol 55: 311-320. http://dx.doi.org/10.1007/sQ0248-
007-9277-3.
Richardson, DHS; Cameron, RP. (2004). Cyanolichens: Their response to pollution and possible
management strategies for their conservation in northeastern North America. Northeast
Nat 11: 1-22. http://dx.doi.org/10.1656/1092-6194(2004^)01 ir0001:CTRTPA12.0.CQ:2.
Richardson, TL; Pincknev, JL; Paerl, HW. (2001). Responses of estuarine phytoplankton
communities to nitrogen form and mixing using microcosm bioassays. Estuaries 24: 828-
839. http://dx.doi.org/10.2307/1353174.
June 2018
R-180
DRAFT: Do Not Cite or Quote

-------
Riddell. J; Jo van. S; Padgett PE; Sweat K. (2011). Tracking lichen community composition
changes due to declining air quality over the last century: The Nash legacy in Southern
California. In ST Bates; F Bungartz; R Lucking; MA Herrera-Campos; A Zambrano
(Eds.), Tracking Lichen Community Composition Changes due to Declining Air Quality
over the Last Century: The Nash Legacy in Southern California (pp. 263-277). Stuttgart,
Germany: Cramer in der Gebr. Borntraeger Verlagsbuchhandlung.
http://www.treesearch.fs.fed.us/pubs/40295.
Riddell J: Nash. TH. Ill: Padgett P. (2008). The effect of HN03 gas on the lichen Ramalina
menziesii. Flora 203: 47-54. http://dx.doi.Org/10.1016/i.flora.2007.10.001.
Riddell J: Padgett PE: Nash. TH. III. (2012). Physiological responses of lichens to factorial
fumigations with nitric acid and ozone. Environ Pollut 170: 202-210.
http://dx.doi.Org/10.1016/i.envpol.2012.06.014.
Riggan. PJ; Lockwood. RN: Jacks. PM; Colver. CG: Weirich. F; DeBano. LF; Brass. JA. (1994).
Effects of fire severity on nitrate mobilization in watersheds subject to chronic
atmospheric deposition. Environ Sci Technol 28: 369-375.
http://dx.doi.org/10.1021/esQ0052a005.
Riggan. PJ: Lockwood. RN: Lopez. EN. (1985). Deposition and processing of airborne nitrogen
pollutants in Mediterranean-type ecosystems of southern California. Environ Sci Technol
19: 781-789. http://dx.doi.org/10.1021 /esOO 139a003.
Riggs. CE; Hobbie. SE; Bach. EM: Hofmockel. KS: Kazanski. CE. (2015). Nitrogen addition
changes grassland soil organic matter decomposition. Biogeochemistry 125: 203-219.
http://dx.doi.org/10.1007/slQ533-015-0123-2.
Rillig. MC. (2004). Arbuscular mycorrhizae and terrestrial ecosystem processes. Ecol Lett 7:
740-754. http://dx.doi.org/10.111 l/i.l461-0248.2004.00620.x.
Rindelaub, JD; Mcavey, KM; Shepson, PB. (2015). The photochemical production of organic
nitrates from a-pinene and loss via acid-dependent particle phase hydrolysis. Atmos
Environ 100: 193-201. http://dx.doi.Org/10.1016/i.atmosenv.2014.l 1.010.
Rizzetto. S: Belvazid. S: Gegout. JC: Nicolas. M; Alard. D; Corcket. E; Gaudio. N: Sverdrup. H;
Probst, A. (2016). Modelling the impact of climate change and atmospheric N deposition
on French forests biodiversity. Environ Pollut 213: 1016-1027.
http://dx.doi.Org/10.1016/i.envpol.2015.12.048.
RMNP Initiative (Rocky Mountain National Park Initiative). (2014). Rocky Mountain National
Park initiative: 2012 nitrogen deposition milestone report. Colorado Department of
Public Health and Environment, U.S. Environmental Protection Agency, National Park
Service. https://www.colorado.gov/pacific/sites/default/files/AP PO RMNP-Initiative-
2013-Milestone-Report.pdf.
RMNP Initiative (Rocky Mountain National Park Initiative). (2015). RMNP Initiative nitrogen
deposition current status as of May 2015. Available online at
https://www.colorado.gov/pacific/sites/default/files/AP-RMNP-status.pdf (accessed
Roberts. KL; Kessler. AJ; Grace. MR: Cook. PLM. (2014). Increased rates of dissimilatory
nitrate reduction to ammonium (DNRA) under oxic conditions in a periodically hypoxic
estuary. Geochim Cosmo Act 133: 313-324. http://dx.doi.Org/10.1016/i.gca.2014.02.042.
June 2018
R-181
DRAFT: Do Not Cite or Quote

-------
Robertson. DM; Saacl. DA. (2013). SPARROW models used to understand nutrient sources in
the Mississippi/Atchafalaya river basin. J Environ Qual 42: 1422-1440.
http://dx.doi.ore/10.2134/ieq2013.02.0066.
Robertson. EK; Roberts. KL; Burdorf. LDW: Cook. P; Thamdrup. B. o. (2016). Dissimilatory
nitrate reduction to ammonium coupled to Fe(II) oxidation in sediments of a periodically
hypoxic estuary. Limnol Oceanogr 61: 365-381. http://dx.doi.org/10.1002/lno. 10220.
Robinson. RB; Barnett. TW: Harwell. GR; Moore. SE; Kulp. M; Schwartz. JS. (2008). pH and
acid anion time trends in different elevation ranges in the Great Smoky Mountains
National Park. J Environ Eng 134: 800-808.
Robinson. RB: Wood. MS: Smoot. JL; Moore. SE. (2004). Parametric modeling of water quality
and sampling strategy in a high-altitude appalachian stream. J Hydrol 287: 62-73.
http ://dx. doi. org/10.1016/i ,i hydrol .2003.09.024.
Robison, AL; Scanlon, TM; Cosby, BJ; Webb, JR; Galloway, JN. (2013). Roles of sulfate
adsorption and base cation supply in controlling the chemical response of streams of
western Virginia to reduced acid deposition. Biogeochemistry 116: 119-130.
http://dx.doi.org/10.1007/slQ533-013-9921-6.
Rochefort, L; Vitt, DH; Bayley, SE. (1990). Growth, production, and decomposition dynamics of
sphagnum under natural and experimentally acidified conditions. Ecology 71: 1986-2000.
http://dx.doi.org/10.2307/1937607.
Rockstrom, J: Steffen, W: Noone, K; Persson, A, sa: Chapin, FS, III: Lambin, EF; Lenton, TM:
Scheffer, M; Folke, C; Schellnhuber, HJ: Nykvist, B; de Wit, CA; Hughes, T; van Per
Leeuw, S: Rodhe, H; Sorlin, S: Snyder, PK; Costanza, R; Svedin, U; Falkenmark, M;
Karlberg, L; Corell, RW: Fabry, VJ; Hansen, J: Walker, B; Liverman, D; Richardson, K;
Crutzen, P; Foley, JA. (2009). A safe operating space for humanity. Nature 461: 472-475.
http://dx.doi.org/10.1038/461472a.
Rodriguez-Cardona, B; Wvmore, AS: Mcdowell, WH. (2016). DOC:N03- ratios and N03-
uptake in forested headwater streams. Jour Geo Res: Biog 121: 205-217.
http://dx.doi.org/10.1002/2015JG0Q3146.
Roelofs, JGM. (1986). The effect of airborne sulphur and nitrogen deposition on aquatic and
terrestrial heathland vegetation. Cell Mol Life Sci 42: 372-377.
http://dx.doi.org/10.1007/BF02118618.
Roem, WJ: Klees, H; Berendse, F. (2002). Effects of nutrient addition and acidification on plant
species diversity and seed germination in heathland. J Appl Ecol 39: 937-948.
http://dx.doi.Org/10.1046/i.1365-2664.2002.00768.x.
Rogers, PC: Moore, KD; Ryel, RJ. (2009). Aspen succession and nitrogen loading: A case for
epiphytic lichens as bioindicators in the Rocky Mountains, USA. J Veg Sci 20: 498-510.
http://dx.doi.org/10.1111/j. 1654-1103.2009.01064.x.
Roig, N: Sierra, J: Rovira, J: Schuhmacher, M; Domingo, JL: Nadal, M. (2013). In vitro tests to
assess toxic effects of airborne PM(10) samples. Correlation with metals and chlorinated
dioxins and furans. Sci Total Environ 443: 791-797.
http: //dx. doi. or g/10.1016/j. scitotenv .2012.11.022.
June 2018
R-182
DRAFT: Do Not Cite or Quote

-------
Rollins. AW; Kiendler-Scharr. A; Fry. JL; Brauers. T; Brown. SS; Porn. HP; Dube. WP; Fuchs.
H; Mensah. A; Mentel. TF; Rohrer. F; Tillmann. R; Wegener. R; Wooldridge. PJ; Cohen.
RC. (2009). Isoprene oxidation by nitrate radical: Alkyl nitrate and secondary organic
aerosol yields. Atmos Chem Phys 9: 6685-6703. http://dx.doi.org/10.5194/acp-9-6685-
2009.
Romanowicz. KJ; Zak. DJ. (2017). Activity of an introduced earthworm (Lumbricus terrestris)
increases under future rates of atmospheric nitrogen deposition in northern temperate
forests. Appl Soil Ecol 120: 206-210. http://dx.doi.Org/10.1016/i.apsoil.2017.08.007.
Romme. WH; Knight. DH. (1981). Fire frequency and subalpine forest succession along a
topographic gradient in Wyoming. Ecology 62: 319-326.
http://dx.doi.org/10.2307/1936706.
Root. HT; Geiser. LH; Fenn. ME; Jovan. S; Hutten. MA; Ahuia. S; Dillman. K; Schirokauer. D;
Berryman. S; McMurrav. JA. (2013). A simple tool for estimating throughfall nitrogen
deposition in forests of western North America using lichens. For Ecol Manage 306: 1-8.
http://dx.doi.Org/10.1016/i.foreco.2013.06.028.
Root. HT; Geiser. LH; Jovan. S; Neitlich. P. (2015). Epiphytic macrolichen indication of air
quality and climate in interior forested mountains of the Pacific Northwest, USA. Ecol
Indicat 53: 95-105. http://dx.doi.Org/10.1016/i.ecolind.2015.01.029.
Roscioli. JR; Zahniser. MS; Nelson. DP; Herndon. SC; Kolb. CE. (2016). New approaches to
measuring sticky molecules: Improvement of instrumental response times using active
passivation. J Phys Chem A 120: 1347-1357. http://dx.doi.org/10.1021/acs.ipca.5b04395.
Rose. JM; Bricker. SB; Ferreira. JG. (2015a). Comparative analysis of modeled nitrogen removal
by shellfish farms. Mar Pollut Bull 91: 185-190.
http://dx.doi.Org/10.1016/i.marpolbul.2014.12.006.
Rose, JM; Bricker, SB; Tedesco, MA; Wikfors, GH. (2014). A role for shellfish aquaculture in
coastal nitrogen management. Environ Sci Technol 48: 2519-2525.
http://dx.doi.org/10.1021/es4041336.
Rose, LA; Elliott, EM; Adams, MB. (2015b). Triple nitrate isotopes indicate differing nitrate
source contributions to streams across a nitrogen saturation gradient. Ecosystems 18:
1209-1223. http://dx.doi.org/10.1007/slQ021-015-9891-8.
Rose, R, ob; Monteith, D, onT; Henrys, P; Smart, S; Wood, C; Morecroft, M; Andrews, C;
Beaumont, D; Benham, S, ue; Bowmaker, V; Corbett, S; Dick, J, an; Dodd, B, ev; Dodd,
N; Flexen, M, el; Mckenna, C; Mcmillan, S; Pallett, D; Rennie, S, ue; Schafer, S; Scott,
T; Sherrin, L; Turner, A; Watson, H. (2016). Evidence for increases in vegetation species
richness across UK Environmental Change Network sites linked to changes in air
pollution and weather patterns. Ecol Indicat 68: 52-62.
http://dx.doi.Org/10.1016/i.ecolind.2016.01.005.
Rosenblum, EB; Voyles, J; Poorten, TJ; Staiich, JE. (2010). The deadly chytrid fungus: a story
of an emerging pathogen [Review], PLoS Pathog 6: el000550.
http://dx.doi.org/10.1371/iournal.ppat.100055Q.
June 2018
R-183
DRAFT: Do Not Cite or Quote

-------
Ross. DS; Lawrence. GB; Fredriksen, G. (2004). Mineralization and nitrification patterns at eight
northeastern USA forested research sites. For Ecol Manage 188: 317-355.
http://dx.doi.ore/10.1016/i.foreco.2003.08.004.
Ross. DS: Wemple. BC. (2011). Soil nitrification in a large forested watershed, Ranch Brook
(Vermont), mirrors patterns in smaller northeastern USA catchments. For Ecol Manage
262: 1084-1093. http://dx.doi.Org/10.1016/i.foreco.2011.06.002.
Rosseland. BO: Skogheim, OK. (1984). A comparative study on salmonid fish species in acid
aluminum-rich water. II Physiological stress and mortality of one- and two-year-old fish
(pp. 186-194). Drottningholm, Sweden: Institute of Freshwater Research.
Rossignol. KL; Paerl. HW: Fear. JM; Braddv. JS. (2011). Nutrients in precipitation and the
phytoplankton responses to enrichment in surface waters of the Albemarle Peninsula,
NC, USA after the establishment of a large-scale chicken egg farm. Hydrobiologia 671:
181-191. http://dx.doi.org/10.1007/sl0750-011-Q715-3.
Rothenberger. MB: Burkholder. JM: Wentworth. TR. (2009). Use of long-term data and
multivariate ordination techniques to identify environmental factors governing estuarine
phytoplankton species dynamics. Limnol Oceanogr 54: 2107-2127.
Rothenberger, MB; Calomeni, AJ. (2016). Complex interactions between nutrient enrichment
and zooplankton in regulating estuarine phytoplankton assemblages: Microcosm
experiments informed by an environmental dataset. Exp Mar Bio Ecol 480: 62-73.
http://dx.doi.Org/10.1016/i.iembe.2016.03.015.
Rothenberger, MB; Swaffield, T; Calomeni, AJ: Cabrey, CD. (2014). Multivariate analysis of
water quality and plankton assemblages in an urban estuary. Estuaries Coasts 37: 695-
711. http://dx.doi.org/10.1007/sl2237-013-9714-Q.
Rousk, J: Brookes, PC: Baath, E. (2009). Contrasting soil pH effects on fungal and bacterial
growth suggest functional redundancy in carbon mineralization. Appl Environ Microbiol
75: 1589-1596. http://dx.doi.org/10.1128/AEM.02775-08.
Roy, PO: Huiibregts, M: Deschenes, L: Margni, M. (2012). Spatially-differentiated atmospheric
source-receptor relationships for nitrogen oxides, sulfur oxides and ammonia emissions at
the global scale for life cycle impact assessment. Atmos Environ 62: 74-81.
http://dx.doi.Org/10.1016/i.atmosenv.2012.07.069.
Rudd, JWM: Kelly, CA: Furutani, A. (1986). The role of sulfate reduction in longterm
accumulation of organic and inorganic sulfur in lake sediments. Limnol Oceanogr 31:
1281-1291. http://dx.doi.Org/10.4319/lo.1986.31.6.1281.
Rudek, J: Paerl, HW: Mallin, MA: Bates, PW. (1991). Seasonal and hydrological control of
phytoplankton nutrient limitation in the lower Neuse River Estuary, North Carolina. Mar
Ecol Prog Ser 75: 133-142. http://dx.doi.org/10.3354/meps075133.
Ruess, RW: Anderson, MP: McFarland, JM: Kielland, K: Olson, K: Taylor, PL. (2013).
Ecosystem-level consequences of symbiont partnerships in an N-fixing shrub from
interior Alaskan floodplains. Ecol Monogr 83: 177-194. http://dx.doi.org/10.1890/12-
0782.1.
June 2018
R-184
DRAFT: Do Not Cite or Quote

-------
Rueth. H; Baron. JS. (2002). Differences in Englemann spruce forest biogeochemistry east and
west of the continental divide in Colorado, USA. Ecosystems 5: 45-57.
http://dx.doi.org/10.1007/sl0021-001-0Q54-8.
Rueth. HM; Baron. JS; Allstott EJ. (2003). Responses of Engelmann spruce forests to nitrogen
fertilization in the Colorado Rocky Mountains. Ecol Appl 13: 664-673.
http://dx.doi.org/10.1890/1051-0761(2003N)013r0664:ROESFT12.0.CQ:2.
Ruhl. HA: Rybicki. NB. (2010). Long-term reductions in anthropogenic nutrients link to
improvements in Chesapeake Bay habitat. ProcNatl Acad Sci USA 107: 16566-16570.
http://dx.doi.org/10.1073/pnas.10035901Q7.
Rtihland, K; Paterson. AM: Smol. JP. (2008). Hemispheric-scale patterns of climate-related
shifts in planktonic diatoms from North American and European lakes. Global Change
Biol 14: 2740-2754. http://dx.doi.Org/10.l 111/i. 1365-2486.2008.01670.X.
Rtihland, KM; Paterson, AM; Smol, JP. (2015). Lake diatom responses to warming: Reviewing
the evidence. J Paleolimnol 54: 1-35. http://dx.doi.org/10.1007/slQ933-015-9837-3.
Rumsev, IC: Walker, JT. (2016). Application of an online ion chromatography-based instrument
for gradient flux measurements of speciated nitrogen and sulfur. Atmospheric
Measurement Techniques Discussions 9: 2581-2592. http://dx.doi.org/10.5194/amt-2016-
4.
Russell, M; Greening, H. (2015). Estimating Benefits in a Recovering Estuary: Tampa Bay,
Florida. Estuaries Coasts 38: S9-S18. http://dx.doi.org/10.1007/sl2237-013-9662-8.
Russow, R; Spott, O: Stange, CF. (2008). Evaluation of nitrate and ammonium as sources of NO
and N20 emissions from black earth soils (Haplic Chernozem) based on N-15 field
experiments. Soil Biol Biochem 40: 380-391.
http://dx.doi.Org/10.1016/i.soilbio.2007.08.020.
Rustad, LE; Campbell, JL; Marion, GM; Norbv, RJ; Mitchell, MJ; Hartley, AE; Cornelissen,
JHC: Gurevitch, J. (2001). A meta-analysis of the response of soil respiration, net
nitrogen mineralization, and aboveground plant growth to experimental ecosystem
warming. Oecologia 126: 543-562. http://dx.doi.org/10.1007/s00442000Q544.
Ruttenberg, BI; Granek, EF. (2011). Bridging the marine-terrestrial disconnect to improve
marine coastal zone science and management. Mar Ecol Prog Ser 434: 203-212.
http://dx.doi.org/10.3354/meps09132.
Ryan, AB; Bover, KE. (2012). Nitrogen further promotes a dominant salt marsh plant in an
increasingly saline environment. Journal of Plant Ecology 5: 429-441.
http://dx.doi.org/10.1093/ipe/rts001.
Ryan, MG; Hubbard, RM; Pongracic, S; Raison, RJ: McMurtrie, RE. (1996). Foliage, fine-root,
woody-tissue and stand respiration in Pinus radiata in relation to nitrogen status. Tree
Physiol 16: 333-343. http://dx.doi.Org/10.1093/treephys/16.3.333.
Rydval, M; Wilson, R. (2012). The impact of industrial S02 pollution on north Bohemia
conifers. Water Air Soil Pollut 223: 5727-5744. http://dx.doi.org/10.1007/sl 1270-012-
1310-6.
June 2018
R-185
DRAFT: Do Not Cite or Quote

-------
Sabaratnam. S; Gupta. G; Mulchi. C. (1988). Nitrogen dioxide effects on photosynthesis in
soybean. J Environ Qual 17: 143-146.
http://dx.doi.org/10.2134/ieql988.00472425001700010Q23x.
Sabo, RD; Nelson. DM: Eshleman, KN. (2016). Episodic, seasonal, and annual export of
atmospheric and microbial nitrate from a temperate forest. Geophys Res Lett 43: 683-
691. http://dx.doi.org/10.1002/2015GLQ66758.
Sadler. K; Lynam. S. (1988). The influence of calcium on aluminium-induced changes in the
growth rate and mortality of brown trout, Salmo trutta L. J Fish Biol 33: 171-179.
http://dx.doi.Org/10.llll/i.1095-8649.1988.tb05460.x.
Saebo. A: Popek. R; Nawrot. B; Hanslin. HM; Gawronska. H; Gawronski. SW. (2012). Plant
species differences in particulate matter accumulation on leaf surfaces. Sci Total Environ
427: 347-354. http://dx.doi.Org/10.1016/i.scitotenv.2012.03.084.
Sahoo, GB; Nover, DM; Reuter, JE; Heyvaert, AC: Riverson, J: Schladow, SG. (2013). Nutrient
and particle load estimates to Lake Tahoe (CA-NV, USA) for total maximum daily load
establishment. Sci Total Environ 444: 579-590.
http://dx.doi.Org/10.1016/i.scitotenv.2012.12.019.
Salisbury, J: Green, M; Hunt, C; Campbell, J. (2008). Coastal acidification by rivers: A threat to
shellfish? Eos 89: 513-528. http://dx.doi.org/10.1029/2008E0500Q01.
SAMAB (Southern Appalachian Man and the Biosphere Cooperative). (1996). The southern
Appalachian assessment: Aquatic technical report. Atlanta, GA: U.S. Dept. of
Agriculture, Forest Service, Southern Region.
http:// samab .org/ saa/reports/aquatic/aquatic.html.
SanClements, MP: Fernandez, IJ; Norton, SA. (2010). Soil chemical and physical properties at
the Bear Brook Watershed in Maine, USA. Environ Monit Assess 171:111-128.
http://dx.doi.org/10.1007/slQ661-010-1531-3.
SanClements, MP: Oelsner, GP; Mcknight, DM: Stoddard, JL; Nelson, SJ. (2012). New insights
into the source of decadal Increases of dissolved organic matter in acid-sensitive lakes of
the Northeastern United States. Environ Sci Technol 46: 3212-3219.
http://dx.doi.org/10.1021/es204321x.
Sander, SP; Abbatt, JPD; Barker, JR; Burkholder, JB; Friedl, RR; Golden, DM: Huie, RE: Kolb,
CE; Kurylo, MJ; Moortgat, GK; Orkin, VL; Wine, PH. (2011). Chemical kinetics and
photochemical data for use in atmospheric studies: Evaluation number 17. (JPL
Publication 10-6). Pasadena, CA: Jet Propulsion Laboratory.
http://ipldataeval.ipl.nasa.gov/pdf/JPL%2010-6%20Final%2015June2011.pdf.
Sanderson, MG: Dentener, FJ; Fiore, AM: Cuvelier, C: Keating, TJ; Zuber, A: Atherton, CS:
Bergmann, DJ; Diehl, T; Dohertv, RM; Duncan, BN: Hess, P; Horowitz, LW: Jacob, DJ:
Jonson, JE; Kaminski, JW; Lupu, A; Mackenzie, IA; Mancini, E; Marmer, E; Park, R;
Pitari, G: Prather, MJ: Pringle, KJ; Schroeder, S: Schultz, MG: Shindell, DT; Szopa, S:
Wild, O: Wind, P. (2008). A multi-model study of the hemispheric transport and
deposition of oxidised nitrogen. Geophys Res Lett 35: L17815.
http://dx.doi.org/10.1029/2008GLQ35389.
June 2018
R-186
DRAFT: Do Not Cite or Quote

-------
Santhi. C; Arnold. JG; White. M; Di Luzio. M; Kantian. N; Norfleet L; Atwood. J; Kellogg. R;
Wang. X; Williams. JR; Gerik. T. (2014). Effects of agricultural conservation practices
on N loads in the Mississippi-Atchafalaya river basin. J Environ Qual 43: 1903-1915.
http://dx.doi.org/10.2134/ieq2013.10.04Q3.
Sanz. MJ: Gries. C: Nash. TH. III. (1992). Dose-response relationships for S02 fumigations in
the lichens Evernia prunastri (L.) Ach. and Ramalina fraxinea (L.) Ach. New Phytol 122:
313-319. http://dx.doi.Org/10.llll/i.1469-8137.1992.tb04236.x.
Sardans. J: Alonso. R; Carnicer. J: Fernandez-Martinez. M; Vivanco. MG: Penuelas. J. (2016).
Factors influencing the foliar elemental composition and stoichiometry in forest trees in
Spain. Perspectives in Plant Ecology, Evolution and Systematics 18: 52-69.
http://dx.doi.Org/10.1016/i.ppees.2016.01.001.
Saros. JE. (2014). Determining critical nitrogen loads to Boreal Lake ecosystems. The response
of phytoplankton at Acadia and Isle Royale National Parks. (National Resources
Technical Report NPS/ACAD/NRTR-2014/862). Fort Collins, CO: U.S. Department of
the Interior, National Park Service.
https://irma.nps.gov/DataStore/Reference/Profile/2209241.
Saros. JE: Clow. DW: Blett. T; Wolfe. AP. (2011). Critical nitrogen deposition loads in high-
elevation lakes of the western US inferred from paleolimnological records. Water Air
Soil Pollut 216: 193-202. http://dx.doi.org/10.1007/sll270-010-Q526-6.
Saros. JE: Interlandi. SJ: Wolfe. AP: Engstrom. DR. (2003). Recent changes in the diatom
community structure of lakes in the Beartooth Mountain Range, USA. Arctic Antarct
Alpine Res 35: 18-23.
Saros. JE: Michel. TJ; Interlandi. SJ: Wolfe. AP. (2005). Resource requirements of Asterionella
formosa and Fragilaria crotonensis in oligotrophic Alpine lakes: Implications for recent
phytoplankton community reorganizations. Can J Fish Aquat Sci 62: 1681-1689.
http://dx.doi.Org/10.l 139/F05-077.
Saros. JE: Rose. KC: Clow. DW: Stephens. VC: Nurse. AB; Arnett. HA: Stone. JR: Williamson.
CE: Wolfe. AP. (2010). Melting Alpine glaciers enrich high-elevation lakes with reactive
nitrogen. Environ Sci Technol 44: 4891-4896. http://dx.doi.org/10.1021/esl00147i.
Saros. JE: Stone. JR: Pederson. GT: Slemmons. KE: Spanbauer. T; Schliep. A: Cahl. D:
Williamson. CE: Engstrom. DR. (2012). Climate-induced changes in lake ecosystem
structure inferred from coupled neo- and paleoecological approaches. Ecology 93: 2155-
2164.
Sauve. S. (2001). Speciation of metals in soils. In HE Allen (Ed.), Bioavailability of metals in
terrestrial ecosystems: Importance of partitioning for bioavailability to invertebrates,
microbes and plants (pp. 7-38). Pensacola, FL: SET AC Press.
Sawicka, K; Rowe, EC: Evans, CD: Monteith, DT; EI Vanguelova, DT; Wade, AJ; J M Clark,
AJ. (2017). Modelling impacts of atmospheric deposition and temperature on long-term
DOC trends. Sci Total Environ 578: 323-336.
http://dx.doi.Org/10.1016/i.scitotenv.2016.10.164.
Saxe, H. (1986). Stomatal-dependent and stomatal-independent uptake of NOx. New Phytol 103:
199-205. http://dx.doi.Org/10.llll/i.1469-8137.1986.tb00608.x.
June 2018
R-187
DRAFT: Do Not Cite or Quote

-------
Scavia. D; Justic. D; Bierman. VJ. (2004). Reducing hypoxia in the Gulf of Mexico: Advice
from three models. Estuaries 27: 419-425. http://dx.doi.org/10.1007/BF028Q3534.
Sea via. D; Kelly. ELA; Hagy. JD. III. (2006). A simple model for forecasting the effects of
nitrogen loads on Chesapeake Bay hypoxia. Estuaries Coasts 29: 674-684.
http://dx.doi.org/10.1007/BF02784292.
Scavia. D; Liu. Y. (2009). Exploring estuarine nutrient susceptibility. Environ Sci Technol 43:
3474-3479. http://dx.doi.org/10.1021/es803401v.
Schaberg. PG: DeHaves. DH: Hawlev. GJ: Murakami. PF; Strimbeck. GR; McNultv. SG.
(2002). Effects of chronic N fertilization on foliar membranes, cold tolerance, and carbon
storage in montane red spruce. Can J For Res 32: 1351-1359.
http ://dx.doi.org/10.113 9/X02-05 9.
Schaberg. PG: Hawlev. GJ: Ravback. SA; Halman. JM; Kosiba. AM. (2014). Inconclusive
evidence of Juniperus virginiana recovery following sulfur pollution reductions [Letter],
Proc Natl Acad Sci USA 111: El. http://dx.doi.org/10.1073/pnas.1320526111.
Schaberg. PG: Minocha. R; Long. S: Halman. JM: Hawlev. GJ: Eagar. C. (2011). Calcium
addition at the Hubbard Brook Experimental Forest increases the capacity for stress
tolerance and carbon capture in red spruce (Picea rubens) trees during the cold season.
Trees Struct Funct 25: 1053-1061. http://dx.doi.org/10.1007/sQ0468-011-0580-8.
Schaberg. PG: Perkins. TP: McNultv. SG. (1997). Effects of chronic low-level N additions on
foliar elemental concentrations, morphology, and gas exchange of mature montane red
spruce. Can J For Res 27: 1622-1629. http://dx.doi.Org/10.l 139/x97-124.
Schartau. AK; Moe. SJ: Sandin. L: Mcfarland. B: Raddum. GG. (2008). Macroinvertebrate
indicators of lake acidification: Analysis of monitoring data from UK, Norway and
Sweden. Aquatic Ecology 42: 293-305. http://dx.doi.org/10.1007/slQ452-008-9186-7.
Scheffe. R: Arnold. J: Davis. C: Dubois. J. -J: Evans. DA: Greaver. T; Herrick. J: Hubbell. B:
Liu. L: Lynch. J: Martin. K: Novak. K: Possiel. N: Rea. A: Reff. A: Smith. JT: Tennant.
G: Wagner. P; Waite. R. (2011). Development of an ambient air quality index. EM
September: 16-22.
Scheffe. RD: Lynch. JA: Reff. A: Kelly. JT: Hubbell. B: Greaver. TL: Smith. JT. (2014). The
aquatic acidification index: A new regulatory metric linking atmospheric and
biogeochemical models to assess potential aquatic ecosystem recovery. Water Air Soil
Pollut 225: 1838. http://dx.doi.org/10.1007/sll270-013-1838-Q.
Scheffe. RD: Solomon. PA: Husar. R: Hanlev. T; Schmidt. M: Koerber. M: Gilrov. M: Hembv.
J: Watkins. N: Papp. M: Rice. J: Tikvart. J: Valentinetti. R. (2009). The National
Ambient Air Monitoring Strategy: Rethinking the Role of National Networks. J Air
Waste Manag Assoc 59: 579-590. http://dx.doi.Org/10.3155/1047-3289.59.5.579.
Schein. A: Courtenav. SC: Crane. CS: Teather. KL: van Den Heuvel. MR. (2012). The role of
submerged aquatic vegetation in structuring the nearshore fish community within an
estuary of the southern gulf of St. Lawrence. Estuaries Coasts 35: 799-810.
http://dx.doi.org/10.1007/sl2237-011-9466-7.
June 2018
R-188
DRAFT: Do Not Cite or Quote

-------
Schein, A; Courtenav, SC; Kidd, KA; Campbell KA; van Den Heuvel, MR. (2013). Food web
structure within an estuary of the southern Gulf of St. Lawrence undergoing
eutrophication. Can J Fish Aquat Sci 70: 1805-1812. http ://dx.doi.org/10.1139/cifas-
2013-0251.
Schelker. J; Sponseller. R; Ring. E; Hogbom, L; Lofgren, S; Laudon. H. (2016). Nitrogen export
from a boreal stream network following forest harvesting: seasonal nitrate removal and
conservative export of organic forms. Biogeosciences 13: 1-12.
http://dx.doi.org/10.5194/bg-13-l-2016.
Schimel. JP; Bennett J. (2004). Nitrogen mineralization: Challenges of a changing paradigm.
Ecology 85: 591-602. http://dx.doi.org/10.1890/03-80Q2.
Schindler. DW. (1974). Eutrophi cation and recovery in experimental lakes: Implications for lake
management. Science 184: 897-899. http://dx.doi.Org/10.l 126/science. 184.4139.897.
Schindler, DW; Armstrong, FAJ; Brunskill, GJ; Holmgren, SK. (1971). Eutrophi cation of Lake
227, Experimental Lakes Area, northwestern Ontario, by addition of phosphate and
nitrate. J Fish Res Board Can 28: 1763-1782. http://dx.doi.Org/10.l 139/f71-261.
Schindler, DW: Mills, KH: Mallev, DF; Findlav, PL: Schearer, JA; Davies, IJ; Turner, MA:
Lindsey, GA; Cruikshank, DR. (1985). Long-term ecosystem stress: Effects of years of
experimental acidification on a small lake. Science 228: 1395-1401.
http://dx.doi.Org/10.l 126/science.228.4706.1395.
Schirokauer, D; Geiser, L; Bytnerowicz, A: Fenn, M; Dillman, K. (2014a). Monitoring air
quality in southeast Alaska national parks and forest: Linking ambient and depositional
pollutants with ecological effects. (NPS/SEAN/NRTR2014/839). Fort Collins, CO: U.S.
National Park Service.
https://science.nature.nps.gov/im/units/sean/AuxRep/AC/AC Schirokauer NRTR 2014-
839.pdf.
Schirokauer, D; Geiser, L; Bytnerowicz, A: Fenn, M; Dillman, K. (2014b). Monitoring air
quality in southeast Alaskas National Parks and Forests: Linking atmospheric pollutants
with ecological effects. (NPS/SEAN/NRTR2014/839). Washington, DC: U.S.
Department of the Interior, National Park Service.
https://science.nature.nps.gov/im/units/sean/AuxRep/AC/AC Schirokauer NRTR 2014-
839.pdf.
Schlegel, H; Amundson, RG: Htittermann, A. (1992). Element distribution in red spruce (Picea
rubens) fine roots; evidence for aluminium toxicity at Whiteface Mountain. Can J For
Res 22: 1132-1138. http://dx.doi.Org/10.l 139/x92-150.
Schleper, C; Nicol, GW. (2010). Ammonia-oxidising archaea—physiology, ecology and
evolution [Review], AdvMicrob Physiol 57: 1-41. http://dx.doi.org/10.1016/B978-Q-12-
381045-8.00001-1.
Schlesinger, WH. (1997). Biogeochemistry: An analysis of global change. In Biogeochemistry:
An Analysis of Global Change (2 ed.). San Diego: Academic Press.
Schlesinger, WH. (2005). Biogeochemistry. In WH Schlesinger (Ed.), Biogeochemistry (1 ed.).
Boston, MA: Elsevier. http://store.elsevier.com/Biogeochemistry/isbn-9780080446424/.
June 2018
R-189
DRAFT: Do Not Cite or Quote

-------
Schlesinger. WH; Pilmanis. AM. (1998). Plant-soil interactions in deserts. Biogeochemistry 42:
169-187. http://dx.doi.Org/10.1023/A:1005939924434.
Schmidt AL; Wvsmyk. JKC: Craig. SE: Lotze. HK. (2012). Regional-scale effects of
eutrophication on ecosystem structure and services of seagrass beds. Limnol Oceanogr
57: 1389-1402. http://dx.doi.Org/10.4319/lo.2012.57.5.1389.
Schmidt MWI; Torn. MS: Abiven. S: Dittmar. T; Guggenberger. G: Janssens. IA: Kleber. M:
Koegel-Knabner. I; Lehmann. J: Manning. DAC: Nannipieri. P; Rasse. DP: Weiner. S:
Trumbore. SE. (2011). Persistence of soil organic matter as an ecosystem property.
Nature 478: 49-56. http://dx.doi.org/10.1038/naturel0386.
Schofield. CL; Driscoll. CT. (1987). Fish species distribution in relation to water quality
gradients in the North Branch of the Moose River Basin. Biogeochemistry 3: 63-85.
http://dx.doi.org/10.1007/BF02185185.
Schofield, CL; Keleher, C. (1996). Comparison of brook trout reproductive success and
recruitment in an acidic Adirondack lake following whole lake liming and watershed
liming. Biogeochemistry 32: 323-337. http://dx.doi.org/10.1007/BFQ2187144.
Schofield. CL: Troinar. JR. (1980). Aluminum toxicity to brook trout (Salvelinus fontinalis) in
acidified waters. In TY Toribara; MW Miller; PE Morrow (Eds.), Polluted Rain (pp. 341-
366). New York, NY: Plenum Press. http://dx.doi.org/10.1007/978-l-4613-306Q-8 18.
Schreck. E; Dappe. V: Sarret. G: Sobanska. S: Nowak. D; Nowak. J: Stefaniak. EA; Magnin. V:
Ranieri. V: Dumat. C. (2014). Foliar or root exposures to smelter particles: Consequences
for lead compartmentalization and speciation in plant leaves. Sci Total Environ 476-477:
667-676. http://dx.doi.Org/10.1016/i.scitotenv.2013.12.089.
Schreck. E; Foucault. Y; Sarret. G: Sobanska. S: Cecillon. L; Castrec-Rouelle. M; Uzu. G:
Dumat. C. (2012). Metal and metalloid foliar uptake by various plant species exposed to
atmospheric industrial fallout: mechanisms involved for lead. Sci Total Environ 427-428:
253-262. http://dx.doi.Org/10.1016/i.scitotenv.2012.03.051.
Schreck. E; Laplanche. C: Le Guedard. M; Bessoule. JJ; Austruv. A: Xiong. T; Foucault. Y;
Dumat. C. (2013). Influence of fine process particles enriched with metals and metalloids
on Lactuca sativa L. leaf fatty acid composition following air and/or soil-plant field
exposure. EnvironPollut 179: 242-249. http://dx.doi.Org/10.1016/i.envpol.2013.04.024.
Schulte-Bisping. H; Beese. F. (2016). N-fluxes and N-turnover in a mixed beech-pine forest
under low N-inputs. European Journal of Forest Research 135: 229-241.
http://dx.doi.org/10.1007/slQ342-015-0931-x.
Schulwitz. SE: Chumchal. MM: Johnson. JA. (2015). Mercury Concentrations in Birds from
Two Atmospherically Contaminated Sites in North Texas, USA. Arch Environ Contam
Toxicol 69: 390-398. http://dx.doi.org/10.1007/sQ0244-015-0189-9.
Schuster. MJ. (2016). Increased rainfall variability and N addition accelerate litter decomposition
in a restored prairie. Oecologia 180: 645-655. http://dx.doi.org/10.1007/sQ0442-015-
3396-1.
June 2018
R-190
DRAFT: Do Not Cite or Quote

-------
Schwab. JJ; Li. Y; Bae, M, -S; Demerjian, KL; Ftou, J; Zhou. X; Jensen. B; Pryor, SC. (2007). A
laboratory intercomparison of real-time gaseous ammonia measurement methods.
Environ Sci Technol 41: 8412-8419. http://dx.doi.org/10.1021/esQ70354r.
Schwartz. NB; Urban. PL: White. PS: Moody. A: Klein. RN. (2016). Vegetation dynamics vary
across topographic and fire severity gradients following prescribed burning in Great
Smoky Mountains National Park. For Ecol Manage 365: 1-11.
http://dx.doi.ore/10.1016/i.foreco.2016.01.027.
Schwarz. GE; Alexander. RB; Smith. RA; Preston. SD. (2011). The regionalization of national-
scale sparrow models for stream nutrients. J Am Water Resour Assoc 47: 1151-1172.
http://dx.doi.Org/10.llll/i.1752-1688.2011.00581.x.
Schwede. D; Zhang. LM; Vet. R; Lear. G. (2011). An intercomparison of the deposition models
used in the CASTNET and CAPMoN networks. Atmos Environ 45: 1337-1346.
http://dx.doi.Org/10.1016/i.atmosenv.2010.l 1.050.
Schwede. DB; Dennis. RL; Bitz. M. (2009). The watershed deposition tool: A tool for
incorporating atmospheric deposition in water-quality analyses(l). J Am Water Resour
Assoc 45: 973-985. http://dx.doi.org/10.111 l/i.l752-1688.2009.00340.x.
Schwede, DB; Lear, GG. (2014a). A novel hybrid approach for estimating total deposition in the
United States. Atmos Environ 92: 207-220.
http://dx.doi.Org/10.1016/i.atmosenv.2014.04.008.
Schwede, DB: Lear, GG. (2014b). A novel hybrid approach for estimating total deposition in the
United States - Supplemental information [Supplemental Data], Atmos Environ 92: 207-
220. http://dx.doi.Org/10.1016/i.atmosenv.2014.04.008.
Scott, JT; Doyle, RD; Prochnow, SJ: White, JD. (2008). Are watershed and lacustrine controls
on planktonicN2 fixation hierarchically structured? Ecol Appl 18: 805-819.
http://dx.doi.Org/10.1890/07-0105.l.
Scott, JT: McCarthy, MJ. (2010). Nitrogen fixation may not balance the nitrogen pool in lakes
over timescales relevant to eutrophication management. Limnol Oceanogr 55: 1265-
1270. http://dx.doi.Org/10.4319/lo.2010.55.3.1265.
Seastedt, TR; Vaccaro, L. (2001). Plant species richness, productivity, and nitrogen and
phosphorus limitations across a snowpack gradient in alpine tundra, Colorado, USA.
Arctic Antarct Alpine Res 33: 100-106. http://dx.doi.org/10.2307/1552283.
Sebestven, SD: Bover, EW: Shanlev, JB; Kendall, C: Doctor, DH; Aiken, GR; Ohte, N. (2008).
Sources, transformations, and hydrological processes that control stream nitrate and
dissolved organic matter concentrations during snowmelt in an upland forest. Water
Resour Res 44. http://dx.doi.org/10.1029/2008WR0Q6983.
Segschneider, H, -J: Htitzen, H; Forstel, H; Rockel, P; Wildt, J. (1993). Uptake of (N02)-N-15
and metabolic transfer of the (N02)-N-15 nitrogen to various nitrogen fractions of
sunflowers. Isotopes Environ Health Stud 29: 51-57.
http://dx.doi.org/10.1080/10256019308Q46135.
June 2018
R-191
DRAFT: Do Not Cite or Quote

-------
Seinfeld. JH; Pandis. SN. (1998). Atmospheric chemistry and physics: From air pollution to
climate change. In Atmospheric Chemistry and Physics: From Air Pollution to Climate
Change. New York: John Wiley & Sons, http://dx.doi.ore/10.1063/1.882420.
Seitzinger. S: Harrison. JA; Bohlke. JK; Bouwman. AF; Lowrance. R; Peterson. B; Tobias. C:
Van Drecht. G. (2006). Denitrification across landscapes and waterscapes: A synthesis.
Ecol Appl 16: 2064-2090. http://dx.doi.org/10.189Q/1051-
0761 (2006)016r2064:DALAWA12.0.CQ:2.
Selvendiran. P; Driscoll. CT; Bushev. JT; Montesdeoca. MR. (2008a). Wetland influence on
mercury fate and transport in a temperate forested watershed. Environ Pollut 154: 46-55.
http://dx.doi.Org/10.1016/i.envpol.2007.12.005.
Selvendiran. P; Driscoll. CT: Montesdeoca. MR: Bushev. JT. (2008b). Inputs, storage, and
transport of total and methyl mercury in two temperate forest wetlands. J Geophys Res
113. http://dx.doi.org/10.1029/2008JG00Q739.
Selvendiran. P; Driscoll. CT: Montesdeoca. MR: Choi. H. -D; Holsen. TM. (2009). Mercury
dynamics and transport in two Adirondack lakes. Limnol Oceanogr 54: 413-427.
http://dx.doi.Org/10.4319/lo.2009.54.2.0413.
Seminoff, JA. (2004). Chelonia mydas. The IUCN Red List of Threatened Species 2004.
Available online at http://www.iucnredlist.Org/details/4615/0 (accessed March 16, 2016).
Serrano. I; Buffam. I; Palm. D; Brannas. E; Laudon. H. (2008). Thresholds for survival of brown
trout during the spring flood acid pulse in streams high in dissolved organic carbon.
Trans Am Fish Soc 137: 1363-1377. http://dx.doi.Org/10.1577/T07-069.l.
Shan. XQ: Wang. ZW: Wang. WS: Zhang. SZ; Wen. B. (2003). Labile rhizosphere soil solution
fraction for prediction of bioavailability of heavy metals and rare earth elements to plants.
Anal Bioanal Chem 375: 400-407. http://dx.doi.org/10.1007/sQ0216-002-1711-2.
Shangguan. Y; Glibert. PM; Alexander. J: Madden. CJ: Murasko. S. (2017). Phytoplankton
assemblage response to changing nutrients in Florida Bay: Results of mesocosm studies.
Exp Mar Bio Ecol 494: 38-53.
Shanlev. JB; Mast. MA: Campbell. DH; Aiken. GR; Krabbenhoft. DP: Hunt. RJ; Walker. JF;
Schuster. PF; Chalmers. A: Aulenbach. BT; Peters. NE; Marvin-Dipasquale. M; Clow.
DW: Shafer. MM. (2008). Comparison of total mercury and methylmercury cycling at
five sites using the small watershed approach. Environ Pollut 154: 143-154.
http://dx.doi.Org/10.1016/i.envpol.2007.12.031.
Shantz. AA; Burkepile. DE. (2014). Context-dependent effects of nutrient loading on the coral-
algal mutualism. Ecology 95: 1995-2005.
Shao, D; Kang, Y; Wu, S; Wong, MH. (2012). Effects of sulfate reducing bacteria and sulfate
concentrations on mercury methylation in freshwater sediments. Sci Total Environ 424:
331-336. http://dx.doi.org/10.1016/i.scitotenv.2011.09.042.
Shao. S: Driscoll. CT: Johnson. CE; Fahev. TJ; Battles. JJ; Blum. JD. (2016). Long-term
responses in soil solution and stream-water chemistry at Hubbard Brook after
experimental addition of wollastonite. Environ Chem 13: 528-540.
http://dx.doi.org/10.1071/EN15113.
June 2018
R-192
DRAFT: Do Not Cite or Quote

-------
Sharpe. PJ; Baldwin. AH. (2013). Wetland plant species richness across estuarine gradients: The
role of environmental factors and the mid-domain effect. Aquat Bot 107: 23-32.
http://dx.doi.ore/10.1016/i.aquabot.2013.01.005.
Shaver. GR; Bret-Harte. SM; Jones. MH; Johnstone. J: Gough. L; Laundre. J: Chapin. FS. III.
(2001). Species composition interacts with fertilizer to control long-term change in tundra
productivity. Ecology 82: 3163-3181. http://dx.doi.org/10.1890/0Q12-
9658(2001)082r3163:SCIWFT12.0.CO;2.
Shaw. GD; Cisneros. R; Schweizer. D; Sickman. JO: Fenn. ME. (2014). Critical loads of acid
deposition for wilderness lakes in the Sierra Nevada (California) estimated by the Steady-
State Water Chemistry model. Water Air Soil Pollut 225: 1804.
http://dx.doi.org/10.1007/sll270-013-18Q4-x.
Sheeslev. RJ; Schauer. JJ; Hemming. JD; Barman. MA: Geis. SW: Tortorelli. JJ. (2004).
Toxicity of ambient atmospheric particulate matter from the Lake Michigan (USA)
airshed to aquatic organisms. Environ Toxicol Chem 23: 133-140.
http://dx.doi.org/10.1897/02-636.
Sheiblev. RW: Enache. M; Swarzenski. PW: Moran. PW: Foreman. JR. (2014). Nitrogen
deposition effects on diatom communities in lakes from three National Parks in
Washington state. Water Air Soil Pollut 225: 1857. http://dx.doi.org/10.1007/sll27Q-
013-1857-x.
Shelton. AO: Francis. TB; Feist. BE: Williams. GD: Lindquist. A: Levin. PS. (2017). Forty years
of seagrass population stability and resilience in an urbanizing estuary. J Ecol 105: 458-
470. http://dx.doi.org/10.1111/1365-2745.12682.
Shen. WJ: Wu. J: Grimm. NB; Hope. D. (2008). Effects of urbanization-induced environmental
changes on ecosystem functioning in the phoenix metropolitan region, USA. Ecosystems
11: 138-155. http://dx.doi.org/10.1007/sl0021-007-9085-Q.
Shephard. MW: Cadv-Pereira. KE; Luo. M; Henze. DK; Pinder. RW: Walker. JT; Rinsland. CP:
Bash. JO: Zhu. L: Payne. VH: Clarisse. L. (2011). TES ammonia retrieval strategy and
global observations of the spatial and seasonal variability of ammonia. Atmos Chem Phys
11: 10743-10763. http://dx.doi.org/10.5194/acp-11 -10743-2011.
Sheppard. LJ: Leith. I. anD: Mizunuma. T: Cape. JN: Crosslev. A: Leeson. S: Sutton. MA: van
Diik. N: Fowler. D. (2011). Dry deposition of ammonia gas drives species change faster
than wet deposition of ammonium ions: Evidence from a long-term field manipulation.
Global Change Biol 17: 3589-3607. http://dx.doi.org/10.1111/i. 1365-2486.2011,02478.x.
Sherman. LA: Baker. LA: Weir. EP: Brezonik. PL. (1994). Sediment pore-water dynamics of
Little Rock Lake, Wisconsin: Geochemical processes and seasonal and spatial variability.
Limnol Oceanogr39: 1155-1171. http://dx.doi.Org/10.4319/lo.1994.39.5.l 155.
Sherwood, ET. (2017). Tampa Bay water quality assessment [Fact Sheet], (Technical Report
#01-17). Tampa Bay, FL: Tampa Bay Estuary Program.
http://www.tbeptech.org/TBEP TECH PUBS/2017/TBEP 01 17 2016 Decision Matri
x Results Update.pdf.
June 2018
R-193
DRAFT: Do Not Cite or Quote

-------
Sherwood. ET; Greening. HS; Janicki. AJ; Karlen. DJ. (2016). Tampa Bay estuary: Monitoring
long-term recovery through regional partnerships. Regional Studies in Marine Science 4:
1-11. http://dx.doi.Org/10.1016/i.rsma.2015.05.005.
Shi. L; Zhang. H; Liu. T. ao: Zhang. W: Shao. Y; Ha. D; Li. Y; Zhang. C: Cai. X: Rao. X: Lin.
Y; Zhou. L; Zhao. P; Ye. Q: Zou. X: Fu. S. (2016a). Consistent effects of canopy vs.
understory nitrogen addition on the soil exchangeable cations and microbial community
in two contrasting forests. Sci Total Environ 553: 349-357.
http://dx.doi.Org/10.1016/i.scitotenv.2016.02.100.
Shi. Y. ao: Sheng. L; Wang. Z; Zhang. X: He. N: Yu. Q. (2016b). Responses of soil enzyme
activity and microbial community compositions to nitrogen addition in bulk and
microaggregate soil in the temperate steppe of Inner Mongolia. Eurasian Soil Science 49:
1149-1160. http://dx.doi.org/10.1134/S1064229316100124.
Shortle. WC: Smith. KT. (1988). Aluminum-induced calcium deficiency syndrome in declining
red spruce. Science 240: 1017-1018. http://dx.doi.Org/10.l 126/science.240.4855.1017.
Shortle. WC: Smith. KT: Minocha. R; Lawrence. GB; David. MB. (1997). Acidic deposition,
cation mobilization, and biochemical indicators of stress in healthy red spruce. J Environ
Qual 26: 871-876. http://dx.doi.org/10.2134/ieal997.00472425002600030Q38x.
Shukurov. N: Kodirov. O; Peitzsch. M; Kersten. M; Pen-Mouratov. S: Steinberger. Y. (2014).
Coupling geochemical, mineralogical and microbiological approaches to assess the health
of contaminated soil around the Almalyk mining and smelter complex, Uzbekistan. Sci
Total Environ 476-477: 447-459. http://dx.doi.Org/10.1016/i.scitotenv.2014.01.031.
Shumchenia. EJ; Pelletier. MC: Cicchetti. G: Davies. S: Pesch. CE; Deacutis. CF; Pryor. M.
(2015). A biological condition gradient model for historical assessment of estuarine
habitat structure. Environ Manage 55: 143-158. http://dx.doi.org/10.1007/sQ0267-014-
0401-0.
Sibold. JS: Veblen. TT; Gonzalez. ME. (2006). Spatial and temporal variation in historic fire
regimes in subalpine forests across the Colorado Front Range in Rocky Mountain
National Park, Colorado, USA. Journal of Biogeography 33: 631-647.
http://dx.doi.org/10.1111/i. 1365-2699.2005.01404.x.
Sickles. JE. II; Hodson. LL. (1999). Retention of sulfur dioxide by nylon filters. Atmos Environ
33: 2423-2426. http://dx.doi.org/10.1016/S1352-2310(98^)00426-9.
Sickles. JE. II; Hodson. LL; Vorburger. LM. (1999). Evaluation of the filter pack for long-
duration sampling of ambient air. Atmos Environ 33: 2187-2202.
http ://dx. doi .org/10.1016/S13 52-2310(98^)00425-7.
Sickles. JE. II; Shadwick. DS; Kilaru. JV; Appel. KW. (2013). "Transference ratios" to predict
total oxidized sulfur and nitrogen deposition - Part II, modeling results. Atmos Environ
77: 1070-1082. http://dx.doi.Org/10.1016/i.atmosenv.2013.05.015.
Sickles. JE; Shadwick. DS. (2013). "Transference ratios" to predict total oxidized sulfur and
nitrogen deposition - Part I, monitoring results. Atmos Environ 77: 1060-1069.
http://dx.doi.Org/10.1016/i.atmosenv.2012.09.072.
June 2018
R-194
DRAFT: Do Not Cite or Quote

-------
Sickman. JO; Levdecker. A; Melack. JM. (2001). Nitrogen mass balances and abiotic controls on
N retention and yield in high-elevation catchments of the Sierra Nevada, California,
United States. Water Resour Res 37: 1445-1461.
http://dx.doi. org/10.1029/2000WR9003 71.
Sickman. JO: Melack. JM: Clow. DW. (2003). Evidence for nutrient enrichment of high-
elevation lakes in the Sierra Nevada, California. Limnology 48: 1885-1892.
http://dx.doi.Org/10.4319/lo.2003.48.5.1885.
Sickman. JO: Melack. JM: Stoddard. JL. (2002). Regional analysis of inorganic nitrogen yield
and retention in high-elevation ecosystems of the Sierra Nevada and Rocky Mountains.
Biogeochemistry 57: 341-374. http://dx.doi.Org/10.1023/A:1016564816701.
Sicolo. M; Tringali. M; Fumagalli. P; Santagostino. A. (2010). Columba livia as a sentinel
species for the assessment of urban air genotoxicity. Arch Environ Contam Toxicol 59:
484-491. http://dx.doi.org/10.1007/sQ0244-010-9488-3.
Sicolo. M; Tringali. M; Orsi. F; Santagostino. A. (2009). Porphyrin pattern and methemoglobin
levels in Columba livia applied to assess toxicological risk by air pollution in urban areas.
Arch Environ Contam Toxicol 57: 732-740. http://dx.doi.org/10.1007/s00244-009-932Q-
0.
Sievering. H. (1999). Nitrogen deposition and carbon sequestration [Letter], Nature 400: 629-
630. http://dx.doi.org/10.1038/23176.
Sievering. H: Fernandez. I; Lee. J: Horn. J: Rustad. L. (2000). Forest canopy uptake of
atmospheric nitrogen deposition at eastern US conifer sites: Carbon storage implications?
Global Biogeochem Cycles 14: 1153-1159. http://dx.doi.org/10.1029/2000GB00125Q.
Sievering. H: Rusch. D; Marquez. L. (1996). Nitric acid, particulate nitrate and ammonium in the
continental free troposphere: nitrogen deposition to an alpine tundra ecosystem. Atmos
Environ 30: 2527-2537. http://dx.doi.org/10.1016/1352-2310(95)00463-7.
Sigal. LL; Nash. TH. III. (1983). Lichen communities on conifers in southern California
mountains: An ecological survey relative to oxidant air pollution. Ecology 64: 1343-
1354. http://dx.doi.org/10.2307/1937489.
Sigtienza. C: Crowley. DE; Allen. EB. (2006). Soil microorganisms of a native shrub and exotic
grasses along a nitrogen deposition gradient in southern California. Appl Soil Ecol 32:
13-26. http://dx.doi.Org/10.1016/i.apsoil.2005.02.015.
Sime. P. (2005). St. Lucie Estuary and Indian River Lagoon conceptual ecological model.
Wetlands 25: 898-907.
Simkin. SM; Allen. EB: Bowman. WD: Clark. CM: Belnap. J: Brooks. ML: Cade. BS: Collins.
SL; Geiser, LH; Gilliam, FS; Jovan, SE; Pardo, LH; Schulz, BK; Stevens, CJ; Suding,
KN: Throop, HL; Waller, DM. (2016). Conditional vulnerability of plant diversity to
atmospheric nitrogen deposition across the United States. Proc Natl Acad Sci USA 113:
4086-4091. http://dx.doi.org/10.1073/pnas.1515241113.
Simkin, SM; Bedford, BL; Weathers, KC. (2013). Phytotoxic sulfide more important than
nutrients for plants within a groundwater-fed wetland. Ecosystems 16: 1118-1129.
http://dx.doi.org/10.1007/slQ021-013-9671-2.
June 2018
R-195
DRAFT: Do Not Cite or Quote

-------
Simon. H; Baker. KR; Phillips. S. (2012). Compilation and interpretation of photochemical
model performance statistics published between 2006 and 2012. Atmos Environ 61: 124-
139. http://dx.doi.Org/10.1016/i.atmosenv.2012.07.012.
Simon. KS: Chadwick. MA: Huryn. AD: Valett. HM. (2010). Stream ecosystem response to
chronic deposition of N and acid at the Bear Brook Watershed, Maine. Environ Monit
Assess 171: 83-92. http://dx.doi.org/10.1007/slQ661-010-1532-2.
Simonin. HA: Kretser. WA: Bath. DW: Olson. M; Gallagher. J. (1993). In situ bioassays of
brook trout (Salvelinus fontinalis) and blacknose dace (Rhinichthys atratulus) in
Adirondack streams affected by episodic acidification. Can J Fish Aquat Sci 50: 902-912.
http://dx.doi.Org/10.l 139/P93-104.
Simpson. LT; Feller. IC: Chapman. SK. (2013). Effects of competition and nutrient enrichemnt
on Avicennia germinans in the salt marsh-mangrove ecotone. Aquat Bot 104: 55-59.
http://dx.doi.Org/10.1016/i.aquabot.2012.09.006.
Singh. NK; Reyes. WM; Bernhardt. ES: Bhattacharya. R; Meyer. JL; Knoepp. JD; Emanuel. RE.
(2016). Hydro-climatological influences on long-term dissolved organic carbon in a
mountain stream of the Southeastern United States. J Environ Qual 45: 1286-1295.
http://dx.doi.org/10.2134/ieq2015.10.0537.
Sinsabaugh. RL; Belnap. J: Rudgers. J: Kuske. CR; Martinez. N: Sandquist. D. (2015). Soil
microbial responses to nitrogen addition in arid ecosystems. FMICB 6: 819.
http://dx.doi.org/10.3389/fmicb.2015.00819.
Sinsabaugh, RL; Shah, JJF. (2011). Ecoenzymatic stoichiometry of recalcitrant organic matter
decomposition: The growth rate hypothesis in reverse. Biogeochemistry 102: 31-43.
http://dx.doi.org/10.1007/slQ533-010-9482-x.
Sirulnik, AG: Allen, EB; Meixner, T; Allen, MF. (2007a). Impacts of anthropogenic N additions
on nitrogen mineralization from plant litter in exotic annual grasslands. Soil Biol
Biochem 39: 24-32. http://dx.doi.Org/10.1016/i.soilbio.2006.04.048.
Sirulnik, AG: Allen, EB: Meixner, T; Fenn, ME: Allen, MF. (2007b). Changes in N cycling and
microbial N with elevated N in exotic annual grasslands of southern California. Appl Soil
Ecol 36: 1-9. http://dx.doi.Org/10.1016/i.apsoil.2006.l 1.004.
Sivaraiah, B: Ruehland, KM: Labai, AL; Paterson, AM: Smol, JP. (2016). Why is the relative
abundance of Asterionella formosa increasing in a Boreal Shield lake as nutrient levels
decline? J Paleolimnol 55: 357-367. http://dx.doi.org/10.1007/slQ933-016-9886-2.
Skeffington, RA: Cosby, BJ: Whitehead, PG. (2016). Long-term predictions of ecosystem
acidification and recovery. Sci Total Environ 568: 381-390.
http://dx.doi.Org/10.1016/i.scitotenv.2016.06.033.
Skogen, KA: Holsinger, KE: Cardon, ZG. (2011). Nitrogen deposition, competition and the
decline of a regionally threatened legume, Desmodium cuspidatum. Oecologia 165: 261-
269. http://dx.doi.org/10.1007/sQ0442-010-1818-7.
Skogheim, OK; Rosseland, BO. (1986). Mortality of smolt of Atlantic salmon, Salmo salar L., at
low levels of aluminium in acidic softwater. Bull Environ Contam Toxicol 37: 258-265.
http://dx.doi.org/10.1007/BF016Q7759.
June 2018
R-196
DRAFT: Do Not Cite or Quote

-------
Slemmons, KEH; Rodgers, ML; Stone. JR; Saros, JE. (2017). Nitrogen subsidies in glacial
meltwaters have altered planktonic diatom communities in lakes of the US Rocky
Mountains for at least a century. Hydrobiologia 800: 129-144.
http://dx.doi.org/10.1007/slQ750-017-3187-2.
Slemmons, KEH: Saros, JE. (2012). Implications of nitrogen-rich glacial meltwater for
phytoplankton diversity and productivity in alpine lakes. Limnol Oceanogr 57: 1651-
1663. http://dx.doi.Org/10.4319/lo.2012.57.6.1651.
Slemmons, KEH: Saros, JE: Simon. K. (2013). The influence of glacial meltwater on alpine
aquatic ecosystems: A review [Review], Environ Sci Process Impacts 15: 1794-1806.
http://dx.doi.org/10.1039/c3em00243h.
Slemmons. KEH: Saros. JE: Stone. JR: McGowan. S: Hess. CT; Cahl. D. (2015). Effects of
glacier meltwater on the algal sedimentary record of an alpine lake in the central US
Rocky Mountains throughout the late Holocene. J Paleolimnol 53: 385-399.
http://dx.doi.org/10.1007/slQ933-015-9829-3.
Small. MJ; Sutton. MC. (1986). A regional pH-alkalinity relationship. Water Res 20: 335-343.
http ://dx. doi .org/10.1016/0043 -13 54(86^)90081 -3.
Smith, DB; Cannon, WF; Woodruff, LG; Solano, F; Kilburn, JE; Fey, PL. (2013). Geochemical
and mineralogical data for soils of the conterminous United States: U.S. Geological
Survey data series 801 (pp. 19). (Data Series 801). Reston, VA: U.S. Department of the
Interior, U.S. Geological Survey, http://pubs.usgs.gov/ds/801/.
Smith, KT; Shortle, WC; Connolly, JH; Minocha, R; Jellison, J. (2009). Calcium fertilization
increases the concentration of calcium in sapwood and calcium oxalate in foliage of red
spruce. Environ Exp Bot 67: 277-283. http://dx.doi.Org/10.1016/i.envexpbot.2009.07.007.
Smith, RA; Schwarz, GE; Alexander, RB. (1997). Regional interpretation of water-quality
monitoring data. Water Resour Res 33: 2781-2798.
http://dx.doi.org/10.1029/97WR02171.
Smith, SM; Lee, KD. (2006). Responses of periphyton to artificial nutrient enrichment in
freshwater kettle ponds of Cape Cod National Aeashore. Hydrobiologia 571: 201-211.
http://dx.doi.org/10.1007/sl0750-006-Q239-4.
Smol, JP; Stoermer, EF. (2010). The diatoms: Applications for the environmental and earth
sciences. In JP Smol; EF Stoermer (Eds.), (2nd ed.). Cambridge, United Kingdom:
Cambridge University Press.
Smolders, AJP; Lamers, LPM; den Hartog, C: Roelofs, JGM. (2003). Mechanisms involved in
the decline of Stratiotes aloides L in the Netherlands: sulphate as a key variable.
Hydrobiologia 506-509: 603-610.
Smyth, AR; Piehler, MF; Grabowski, JH. (2015). Habitat context influences nitrogen removal by
restored oyster reefs. J Appl Ecol 52: 716-725. http://dx.doi.org/10. Ill 1/1365-
2664.12435.
Smyth, AR; Thompson, SP; Siporin, KN; Gardner, WS; McCarthy, MJ: Piehler, MF. (2013).
Assessing nitrogen dynamics throughout the estuarine landscape. Estuaries Coasts 36:
44-55. http://dx.doi.org/10.1007/sl2237-012-9554-3.
June 2018
R-197
DRAFT: Do Not Cite or Quote

-------
Spares. CRF. S; Siqueira. JO. (2008). Mycorrhiza and phosphate protection of tropical grass
species against heavy metal toxicity in multi-contaminated soil. Biol Fertil Soils 44: 833-
841. http://dx.doi.org/10.1007/s00374-007-Q265-z.
Sobota. DJ: Compton. JE; Harrison. JA. (2013). Reactive nitrogen inputs to U.S. lands and
waterways: How certain are we about sources and fluxes? Front Ecol Environ 11: 82-90.
http://dx.doi.Org/10.1890/l 10216.
Sobota. DJ: Compton. JE: McCrackin. ML: Singh. S. (2015). Cost of reactive nitrogen release
from human activities to the environment in the United States. Environ Res Lett 10.
http://dx.doi.Org/10.1088/1748-9326/10/2/025006.
Sobota. DJ: Harrison. JA: Dahlgren. RA. (2009). Influences of climate, hydrology, and land use
on input and export of nitrogen in California watersheds. Biogeochemistry 94: 43-62.
http://dx.doi.org/10.1007/slQ533-009-9307-v.
Soininen, J: Jamoneau, A; Rosebery, J: Passy, SI. (2016). Global patterns and drivers of species
and trait composition in diatoms. Global Ecology and Biogeography 25: 940-950.
http://dx.doi.org/10. Ill 1/geb. 12452.
Solberg. S: Dobbertin. M: Reinds. G: Lange. H: Andreassen. K: Garcia Fernandez. P;
Hildingsson, A; de Vries, W. (2009). Analyses of the impact of changes in atmospheric
deposition and climate on forest growth in European monitoring plots: A stand growth
approach. For Ecol Manage 258: 1735-1750.
http://dx.doi.Org/10.1016/i.foreco.2008.09.057.
Solomon, D; Lehmann, J: Tekalign, M; Fritzsche, F; Zech, W. (2001). Sulfur fractions in
particle-size separates of the sub-humid Ethiopian highlands as influenced by land use
changes. Geoderma 102: 41-59. http://dx.doi.org/10.1016/S0016-7061(00)00103-8.
Solomon. PA: Crumpler. D; Flanagan. JB; Javantv. RK; Rickman. EE: McDade. CE. (2014).
U.S. national PM2.5 Chemical Speciation Monitoring Networks-CSN and IMPROVE:
Description of networks. J Air Waste Manag Assoc 64: 1410-1438.
http://dx.doi.org/10.1080/10962247.2014.9569Q4.
Solomon. PA: Larson. SM: Fall. T; Cass. GR. (1988). Basinwide nitric acid and related species
concentrations observed during the claremont nitrogen species comparison study. Atmos
Environ 22: 1587-1594. http://dx.doi.org/10.1016/0004-6981(88N)90386-l.
Song. M. -H: Yu. F. -H. (2015). Reduced compensatory effects explain the nitrogen-mediated
reduction in stability of an alpine meadow on the Tibetan Plateau. New Phytol 207: 70-
77. http://dx.doi.org/10. Ill 1/nph. 13329.
Song. Y; Song. C: Meng. H: Swarzenski. CM: Wang. X: Tan. W. (2017). Nitrogen additions
affect litter quality and soil biochemical properties in a peatland of Northeast China. Ecol
Eng 100: 175-185. http://dx.doi.Org/10.1016/i.ecoleng.2016.12.025.
Soranno. PA: Bissell. EG: Cheruvelil. KS: Christel. ST: Collins. SM: Fergus. CE. mi: Filstrup.
CT: Lapierre. JF: Lottig. NR: Oliver. SK: Scott. CE: Smith. NJ: Stopvak. S: Yuan. S:
Bremigan. MT: Downing. JA: Gries. C: Henry. EN: Skaff. NK: Stanley. EH: Stow. CA:
Tan. PN: Wagner. T; Webster. KE. (2015). Building a multi-scaled geospatial temporal
ecology database from disparate data sources: fostering open science and data reuse. 4.
http://dx.doi.org/10.1186/sl3742-015-0067-4.
June 2018
R-198
DRAFT: Do Not Cite or Quote

-------
Sorensen, PO: Templer, PH; Finzi, AC. (2016). Contrasting effects of winter snowpack and soil
frost on growing season microbial biomass and enzyme activity in two mixed-hardwood
forests. Biogeochemistry 128: 141-154. http://dx.doi.ore/10.1007/sl0533-016-0199-3.
Soule, PT. (2011). Changing climate, atmospheric composition, and radial tree growth in a
spruce-fir ecosystem on Grandfather Mountain, North Carolina. Natural Areas Journal
31: 65-74. http://dx.doi.org/10.3375/043.031.01Q8.
Soulsbv. C: Turnbull. D; Langan, SJ: Owen. R; Hirst D. (1995). Long-term trends in stream
chemistry and biology in northeast Scotland: Evidence for recovery. Water Air Soil
Pollut 85: 689-694. http://dx.doi.org/10.1007/BF004769Q9.
Souri, AH: Choi. Y; Jeon, W: Li. X: Pan. S: Diao, L; Westenbarger, DA. (2016). Constraining
NOx emissions using satellite N02 measurements during 2013 DISCOVER-AQ Texas
campaign. Atmos Environ 131: 371-381.
http://dx.doi.Org/10.1016/i.atmosenv.2016.02.020.
Southon. GE; Field. C: Caporn. SJM; Britton. AJ; Power. SA. (2013). Nitrogen deposition
reduces plant diversity and alters ecosystem functioning: Field-scale evidence from a
nationwide survey of UK heathlands. PLoS ONE 8: e59031.
http://dx.doi.org/10.1371/iournal.pone.0059031.
Sowers. AA; Brush. GS. (2014). A paleoecological history of the late precolonial and
postcolonial Mesohaline Chesapeake Bay food web. Estuaries Coasts 37: 1506-1515.
http ://dx. doi .org/10.1007/s 1223 7-014-9781 -x.
Spang, A; Hatzenpichler, R; Brochier-Armanet, C; Rattei, T; Tischler, P; Spieck, E, va; Streit,
W: Stahl. DA: Wagner. M; Schleper. C. (2010). Distinct gene set in two different
lineages of ammonia-oxidizing archaea supports the phylum Thaumarchaeota. Trends
Microbiol 18: 331-340. http://dx.doi.Org/10.1016/i.tim.2010.06.003.
Sparks, JP. (2009). Ecological ramifications of the direct foliar uptake of nitrogen [Review],
Oecologia 159: 1-13. http://dx.doi.org/10.1007/sQ0442-008-1188-6.
Spaulding, SA: Otu, MK; Wolfe, AP; Baron, JS. (2015). Paleolimnological records of nitrogen
deposition in shallow, high-elevation lakes of Grand Teton National Park, Wyoming,
USA. Arctic Antarct Alpine Res 47: 703-717. http://dx.doi.org/10.1657/AAAR0015-0Q8.
Spranger, T; Maximilian, P; Hettelingh, J, -P; Slootweg, J. (2004). Manual on methodologies and
criteria for modeling and mapping critical loads and levels and air pollution effects, risks,
and trends. In Manual on Methodologies and Criteria for Modeling and Mapping Critical
Loads and Levels and Air Pollution Effects, Risks, and Trends. Berlin, Germany: Federal
Environmental Agency. hhttp://www.icpmapping.org/Latest update Mapping Manual.
Sridevi, G: Minocha, R; Turlapati, SA: Goldfarb, KC: Brodie, EL: Tisa, LS: Minocha, SC.
(2012). Soil bacterial communities of a calcium-supplemented and a reference watershed
at the Hubbard Brook Experimental Forest (HBEF), New Hampshire, USA. FEMS
Microbiol Ecol 79: 728-740. http://dx.doi.Org/10.llll/i.1574-6941.2011.01258.x.
St Louis, VL; Rudd, JWM; Kelly, CA; Hall, BP: Rolfhus, KR; Scott, KJ; Lindberg, SE; Dong,
W. (2001). Importance of the forest canopy to fluxes of methyl mercury and total
mercury to boreal ecosystems. Environ Sci Technol 35: 3089-3098.
http://dx.doi.org/10.1021/es001924p.
June 2018
R-199
DRAFT: Do Not Cite or Quote

-------
St. Louis. VL; Rudd, JWM; Kelly. CA; Beatv, KG; Flett RJ: Roulet NT. (1996). Production and
loss of methylmercury and loss of total mercury from boreal forest catchments containing
different types of wetlands. Environ Sci Technol 30: 2719-2729.
http://dx.doi.org/10.1021/es95Q856h.
St.Clair. SB: Lynch. JP. (2005). Element accumulation patterns of deciduous and evergreen tree
seedlings on acid soils: Implications for sensitivity to manganese toxicity. Tree Physiol
25: 85-92. http://dx.doi.Org/10.1093/treephys/25.l.85.
St.Clair. SB: Sharpe. WE: Lynch. JP. (2008). Key interactions between nutrient limitation and
climatic factors in temperate forests: A synthesis of the sugar maple literature. Can J For
Res 38: 401-414. http://dx.doi.org/10.1139/X07-161.
St.Louis. VL: Rudd. JWM: Kelly. CA: Bodalv. RA; Paterson. MJ; Beatv. KG: Hesslein. RH;
Heves. A; Maiewski. AR. (2004). The rise and fall of mercury methylation in an
experimental reservoir. Environ Sci Technol 38: 1348-1358.
http://dx.doi. org/10.1021 /es034424f.
STAC (Chesapeake Bay Program Scientific and Technical Advisory Committee). (2013).
Evaluation of the use of shellfish as a method of nutrient reduction in the Chesapeake
Bay (pp. 1-65). (STAC Publ. #13-005). Edgewater, MD: Scientific and Technical
Advisory Committee, http://www.chesapeake.org/pubs/307 Luckenbach2013.pdf.
Stamenkovic. J; Gustin. MS. (2009). Nonstomatal versus stomatal uptake of atmospheric
mercury. Environ Sci Technol 43: 1367-1372. http://dx.doi.org/10.1021/es801583a.
Stark, LR; Brinda, JC; McLetchie, DN. (2011). Effects of increased summer precipitation and N
deposition on Mojave Desert populations of the biological crust moss Syntrichia
caninervis. J Arid Environ 75: 457-463. http://dx.doi.Org/10.1016/i.iaridenv.2010.12.015.
Statham. PJ. (2012). Nutrients in estuaries - an overview and the potential impacts of climate
change [Review], Sci Total Environ 434: 213-227.
http://dx.doi.Org/10.1016/i.scitotenv.2011.09.088.
Stauffer. RE. (1990). Granite weathering and the sensitivity of alpine lakes to acid deposition.
Limnology 35: 1112-1134. http://dx.doi.org/10.4319/lo. 1990.35.5.1112.
Stauffer. RE; Wittchen. BP. (1991). Effects of silicate weathering on water chemistry in
forested, upland, felsic terrane of the USA. Geochim Cosmo Act 55: 3253-3271.
http://dx.doi. org/10.1016/0016-703 7(9 n90487-P.
Stefanowicz. AM; Niklinska. M; Laskowski. R. (2009). Pollution-induced tolerance of soil
bacterial communities in meadow and forest ecosystems polluted with heavy metals.
European Journal of Soil Biology 45: 363-369.
http://dx.doi.Org/10.1016/i.eisobi.2009.05.005.
Steffan. RJ; Korthals. ET; Winfrey. MR. (1988). Effects of acidification on mercury methylation,
demethylation, and volatilization in sediments from an acid-susceptible lake. Appl
Environ Microbiol 54: 2003-2009.
Steiner, CF; Long, ZT; Krumins, JA; Morin, PJ. (2005). Temporal stability of aquatic food webs:
Partitioning the effects of species diversity, species composition and enrichment. Ecol
Lett 8: 819-828. http://dx.doi.Org/10.l 111/i. 1461-0248.2005.00785.x.
June 2018
R-200
DRAFT: Do Not Cite or Quote

-------
Steinmuller. HE; Graham. SA; White. J. R.; Mckee. M; Mendelssohn. IA. (2016). A decadal-
scale nutrient loading study in a coastal wetland: Impacts on soil microbial processes.
Ecol Eng 97: 58-63. http://dx.doi.ore/10.1016/j.ecoleng.2016.07.008.
Stephenson. K; Aultman. S: Metcalfe. T; Miller. A. (2010). An evaluation of nutrient nonpoint
offset trading in Virginia: A role for agricultural nonpoint sources? Water Resour Res 46:
W04519. http://dx.doi.org/10.1029/2009WR0Q8228.
Sterner. RW. (2008). On the phosphorus limitation paradigm for lakes. Int Rev Hydrobiol 93:
433-445. http://dx.doi.org/10.1002/iroh.200811Q68.
Stets, EG; Kelly, VJ; Crawford, CG. (2014). Long-term trends in alkalinity in large rivers of the
conterminous US in relation to acidification, agriculture, and hydrologic modification.
Sci Total Environ 488: 280-289. http://dx.doi.Org/10.1016/i.scitotenv.2014.04.054.
Stevens, C; Dupre, C; Gaudnik, C; Dorland, E; Pise, N; Gowing, D; Bleeker, A; Alard, D;
Bobbink, R; Fowler, D; Vandvik, V; Corcket, E; Mountford, JO; Aarrestad, PA; Muller,
S; Diekmann, M. (201 la). Changes in species composition of European acid grasslands
observed along a gradient of nitrogen deposition. J Veg Sci 22: 207-215.
http://dx.doi.org/10.1111/i. 1654-1103.2010.01254.x.
Stevens, CJ. (2016). How long do ecosystems take to recover from atmospheric nitrogen
deposition? Biol Conserv 200: 160-167. http://dx.doi.Org/10.1016/i.biocon.2016.06.005.
Stevens, CJ; Pise, NB; Mountford, JO; Gowing, DJ. (2004). Impact of nitrogen deposition on the
species richness of grasslands. Science 303: 1876-1879.
http://dx.doi.Org/10.l 126/science. 1094678.
Stevens, CJ; Dupre, C; Dorland, E; Gaudnik, C; Gowing, DJG; Bleeker, A; Diekmann, M; Alard,
D; Bobbink, R; Fowler, D; Corcket, E; Mountford, JO; Vandvik, V; Aarrestad, PA;
Muller, S; Pise, NB. (2010a). Nitrogen deposition threatens species richness of
grasslands across Europe. Environ Pollut 158: 2940-2945.
http://dx.doi.Org/10.1016/i.envpol.2010.06.006.
Stevens, CJ; Dupre, C; Dorland, E; Gaudnik, C; Gowing, DJG; Bleeker, A; Diekmann, M; Alard,
D; Bobbink, R; Fowler, D; Corcket, E; Mountford, JO; Vandvik, V; Aarrestad, PA;
Muller, S; Pise, NB. (201 lb). The impact of nitrogen deposition on acid grasslands in the
Atlantic region of Europe. Environ Pollut 159: 2243-2250.
http://dx.doi.Org/10.1016/i.envpol.2010.l 1.026.
Stevens, CJ; Lind, EM; Hautier, Y; Harpole, WS; Borer, ET; Hobbie, S; Seabloom, EW;
Ladwig, L; Bakker, JP; Chu, C; Collins, S; Pavies, KF; Firn, J; Hillebrand, H; La Pierre,
KJ; MacPougall, A; Melbourne, B; McCullev, RL; Morgan, J; Orrock, JL; Prober, SM;
Risch, AC; Schuetz, M; Wragg, PP. (2015). Anthropogenic nitrogen deposition predicts
local grassland primary production worldwide. Ecology 96: 1459-1465.
http://dx.doi.Org/10.1890/14-1902.l.
Stevens, CJ; Mountford, JO; Gowing, PJG; Bardgett, RP. (2012). Pifferences in yield,
Ellenberg N value, tissue chemistry and soil chemistry 15 years after the cessation of
nitrogen addition. Plant Soil 357: 309-319. http://dx.doi.org/10.1007/sl 1104-012-1160-4.
June 2018
R-201
DRAFT: Do Not Cite or Quote

-------
Stevens. CJ; Payne. RJ; Kimberlev, A; Smart. SM. (2016). How will the semi-natural vegetation
of the UK have changed by 2030 given likely changes in nitrogen deposition? Environ
Pollut 208: 879-889. http://dx.doi.Org/10.1016/i.envpol.2015.09.013.
Stevens. CJ: Thompson. K; Grime. JP; Long. CJ: Gowing, DJG. (2010b). Contribution of
acidification and eutrophication to declines in species richness of calcifuge grasslands
along a gradient of atmospheric nitrogen deposition. Funct Ecol 24: 478-484.
http://dx.doi.org/10.1111/i. 1365-2435.2009.01663 ,x.
Stief. P. (2013). Stimulation of microbial nitrogen cycling in aquatic ecosystems by benthic
macrofauna: Mechanisms and environmental implications. Biogeosciences 10: 7829-
7846. http://dx.doi.org/10.5194/bg-10-7829-2013.
Stock. ZS: Russo. MR: Butler. TM; Archibald. AT: Lawrence. MG: Telford. PJ: Abraham. NL;
Pyle. JA. (2013). Modelling the impact of megacities on local, regional and global
tropospheric ozone and the deposition of nitrogen species. Atmos Chem Phys 13: 12215-
12231. http://dx.doi.org/10.5194/acp-13-12215-2013.
Stockdale. A: Tipping. E: Fiellheim. A: Garmo. 0A: Hildrew. AG: Lofts. S: Monteith. DT:
Ormerod. SJ: Shilland. EM. (2011). Recovery of macroinvertebrate species richness in
acidified upland waters assessed with a field toxicity model. Ecol Indicat 37: 341-350.
http://dx.doi.Org/10.1016/i.ecolind.2011.11.002.
Stoddard. JL. (1995). Episodic acidification during snowmelt of high elevation lakes in the Sierra
Nevada mountains of California. Water Air Soil Pollut 85: 353-358.
http://dx.doi.org/10.1007/BF0Q476854.
Stoddard. JL. (1994). Long-term changes in watershed retention of nitrogen: Its causes and
aquatic consequences. In LA Baker (Ed.), Environmental Chemistry of Lakes and
Reservoirs (Advances in chemistry series no. 237 ed., pp. 223-284). Washington, DC:
American Chemical Society, http://dx.doi.org/10.1021/ba-1994-0237.ch008.
Stoddard. JL: Driscoll. CT: Kahl. JS: Kellogg. JH. (1998). A regional analysis of lake
acidification trends for the northeastern US, 1982-1994. Environ Monit Assess 51: 399-
413. http://dx.doi.Org/10.1023/A:1005988732077.
Stoddard, JL; Urquhart, NS; Newell, AD; Kugler, D. (1996). The Temporally Integrated
Monitoring of Ecosystems (TIME) project design: 2. Detection of regional acidification
trends. Water Resour Res 32: 2529-2538. http://dx.doi.org/10.1029/96WR009Q2.
Stoddard, JL: Van Sickle, J: Herlihy, AT: Brahnev, J: Paulsen, S: Peck, DV: Mitchell, R:
Pollard, AI. (2016). Continental-Scale Increase in Lake and Stream Phosphorus: Are
Oligotrophic Systems Disappearing in the United States? Environ Sci Technol 50: 3409-
3415. http://dx.doi.org/10.1021/acs.est.5b05950.
Stoermer, EF: Smol, JP. (1999). The diatoms: Applications for the environmental and earth
sciences. In EF Stoermer; JP Smol (Eds.). Cambridge, UK: Cambridge University Press.
Stoklasa, J. (1923). Die beschadigungen der vegetation durch Rauchgase und fabriksexhalation
[Damage to vegetation from flue-gas and factory emissions]. In Die Beschadigungen der
Vegetation durch Rauchgase und Fabriksexhalation [Damage to Vegetation from Flue-
Gas and Factory Emissions], Berlin, Germany: Urban and Schwarzenberg.
June 2018
R-202
DRAFT: Do Not Cite or Quote

-------
Stolte. W; McCollin. T; Noordeloos. A; Riegman. R. (1994). Effects of nitrogen source on the
size distribution within marine phytoplankton populations. Exp Mar Bio Ecol 184: 83-97.
http ://dx. doi .org/10.1016/0022-0981 (94)90167-8.
Stone. JJ; McCutcheon. CM; Stetler. LP; Chipps. SR. (2011). Interrelationships between fish
tissue mercury concentrations and water quality for South Dakota natural lakes and
impoundments. Water Air Soil Pollut 222: 337-349. http://dx.doi.org/10.1007/sl 1270-
011-0828-3.
Stone. JR; Saros. JE; Pederson. GT. (2016). Coherent late-Holocene climate-driven shifts in the
structure of three Rocky Mountain lakes. Holocene 26: 1103-1111.
http://dx.doi.org/10.1177/0959683616632886.
Strengbom. J: Nordin. A. (2008). Commercial forest fertilization causes long-term residual
effects in ground vegetation of boreal forests. For Ecol Manage 256: 2175-2181.
http://dx.doi.Org/10.1016/i.foreco.2008.08.009.
Strengbom. J: Nordin. A: Nasholm. T; Ericson. L. (2001). Slow recovery of boreal forest
ecosystem following decreased nitrogen input. Funct Ecol 15: 451-457.
http://dx.doi.Org/10.1046/i.0269-8463.2001.00538.x.
Strickman, RJ; Fulthorpe, RR; Coleman Wasik, JK; Engstrom, PR; Mitchell, CP. (2016).
Experimental sulfate amendment alters peatland bacterial community structure. Sci Total
Environ 566-567: 1289-1296. http://dx.doi.Org/10.1016/i.scitotenv.2016.05.189.
Strock. KE; Nelson. SJ: Kahl. JS: Saros. JE: McDowell. WH. (2014). Decadal trends reveal
recent acceleration in the rate of recovery from acidification in the Northeastern US.
Environ Sci Technol 48: 4681-4689. http://dx.doi.org/10.1021/es404772n.
Strock. KE: Saros. JE: Nelson. SJ: Birkel. SD; Kahl. JS: Mcdowell. WH. (2016). Extreme
weather years drive episodic changes in lake chemistry: implications for recovery from
sulfate deposition and long-term trends in dissolved organic carbon. Biogeochemistry
127: 353-365. http://dx.doi.org/10.1007/slQ533-016-0185-9.
Stuart. KR; Drawbridge. M. (2011). The effect of light intensity and green water on survival and
growth of cultured larval California yellowtail (Seriola lalandi). Aquaculture 321: 152-
156. http://dx.doi.org/10.1016/j.aquaculture.2011.08.023.
Sturdivant. SK; Brush. MJ; Diaz. RJ. (2013). Modeling the effect of hypoxia on macrobenthos
production in the lower Rappahannock River, Chesapeake Bay, USA. PLoS ONE 8:
e84140. http://dx.doi.org/10.1371/iournal.pone.0084140.
Sturdivant. SK: Diaz. RJ: Llanso. R; Dauer. DM. (2014). Relationship between hypoxia and
macrobenthic production in Chesapeake Bay. Estuaries Coasts 37: 1219-1232.
http://dx.doi.org/10.1007/sl2237-013-9763-4.
Stursova. M; Crenshaw. CL; Sinsabaugh. RL. (2006). Microbial responses to long-term N
deposition in a semiarid grassland. Microb Ecol 51: 90-98.
http://dx.doi.org/10.1007/sQ0248-005-5156-v.
June 2018
R-203
DRAFT: Do Not Cite or Quote

-------
Suding. KN; Collins. SL; Gough. L; Clark. C; Cleland. EE; Gross. KL; Milchunas. DG;
Pennings. S. (2005). Functional- and abundance-based mechanisms explain diversity loss
due toN fertilization. ProcNatl Acad Sci USA 102: 4387-4392.
http://dx.doi.org/10.1073/pnas.04086481Q2.
Sukhdev. P; Wittmer. H; Miller. D. (2014). The economics of ecosystems and biodiversity
(TEEB): Challenges and responses. In The Economics of Ecosystems and Biodiversity
(TEEB): Challenges and Responses (pp. 135-150). Oxford, UK: Oxford University
Presss. http://dx.doi.Org/10.1093/acprof:oso/9780199676880.003.0007.
Sullivan. PF; Sommerkorn. M; Rueth. HM; Nadelhoffer. KJ; Shaver. GR; Welker. JM. (2007a).
Climate and species affect fine root production with long-term fertilization in acidic
tussock tundra near Toolik Lake, Alaska. Oecologia 153: 643-652.
http://dx.doi.org/10.1007/s00442-007-Q753-8.
Sullivan. TJ. (2000). Aquatic effects of acidic deposition. In Aquatic Effects of Acidic
Deposition (1 ed.). Boca Raton, FL: CRC Press.
Sullivan. TJ. (2012). Combining ecosystem service and critical load concepts for resource
management and public policy. 4: 905-913. http://dx.doi.org/10.3390/w40409Q5.
Sullivan, TJ. (2015). Air pollutant deposition and effects on natural resources in New York State.
In Air Pollutant Deposition and Effects on Natural Resources in New York State. Ithaca,
NY: Comstock Publishing Associates. http://www.istor.Org/stable/10.7591/i.cttl5hvshf.
Sullivan. TJ. (2017). Air pollution and its impacts on U.S. national parks. Boca Raton, FL: CRC
Press. https://www.crcpress.com/Air-Pollution-and-Its-Impacts-on-US-National-
Parks/Sullivan/p/book/9781498765176.
Sullivan. TJ: Charles. DF. (1994). The feasibility and utility of a paleolimnology/paleoclimate
data cooperative for North America. J Paleolimnol 10: 265-273.
http://dx.doi.org/10.1007/BF00684Q36.
Sullivan. TJ: Charles. DF: Smol. JP; Cumming. BF; Selle. AR; Thomas. PR: Bernert. JA; Dixit.
SS. (1990). Quantification of changes in lakewater chemistry in response to acidic
deposition. Nature 345: 54-58. http://dx.doi.org/10.1038/345054a0.
Sullivan. TJ: Cosby. BJ; Driscoll. CT; McDonnell. TC: Herlihy. AT. (201 la). Target loads of
atmospheric sulfur deposition protect terrestrial resources in the Adirondack Mountains,
New York against biological impacts caused by soil acidification. Journal of
Environmental Studies and Sciences 1: 301-314. http://dx.doi.org/10.10Q7/sl3412-011-
0062-8.
Sullivan. TJ: Cosby. BJ: Driscoll. CT: McDonnell. TC: Herlihy. AT: Burns. DA. (2012a). Target
loads of atmospheric sulfur and nitrogen deposition for protection of acid sensitive
aquatic resources in the Adirondack Mountains, New York. Water Resour Res 48:
W01547. http://dx.doi.org/10.1029/201 lwrOl 1171.
Sullivan. TJ: Cosby. BJ: Herlihy. AT: Webb. JR: Bulger. AJ: Snyder. KU: Brewer. PF: Gilbert.
EH; Moore. PL. (2004). Regional model projections of future effects of sulfur and
nitrogen deposition on streams in the southern Appalachian Mountains. Water Resour
Res 40: W02101. http://dx.doi.org/10.1029/2003WROO1998.
June 2018
R-204
DRAFT: Do Not Cite or Quote

-------
Sullivan. TJ; Cosby. BJ; Jackson. WA. (201 lb). Target loads of atmospheric sulfur deposition
for the protection and recovery of acid-sensitive streams in the Southern Blue Ridge
Province. J Environ Manage 92: 2953-2960.
http://dx.doi.Org/10.1016/i.ienvman.2011.07.014.
Sullivan. TJ: Cosby. BJ: Jackson. WA: Snyder. K; Herlihy. AT. (201 lc). Acidification and
prognosis for future recovery of acid-sensitive streams in the Southern Blue Ridge
province. Water Air Soil Pollut 219: 11-16. http://dx.doi.ore/10.1007/sl 1270-010-0680-
x.
Sullivan. TJ: Cosby. BJ: Laurence. JA: Dennis. RL: Savig. K: Webb. JR: Bulger. AJ: Scruggs.
M: Gordon. C: Rav. J: Lee. H: Hogsett. WE: Wayne. H: Miller. D: Kern. JS. (2003).
Assessment of air quality and related values in Shenandoah National Park.
(NPS/NERCHAL/NRTR-03/090). Philadelphia, PA: Natural Resource Stewardship and
Science, Northeast Region, National Park Service, U.S. Department of the Interior.
https://www.nature.nps.gov/air/Pubs/pdf/SHEN Assess Sullivan2003.pdf.
Sullivan, TJ: Cosby, BJ: McDonnell, TC: Porter, EM: Blett, T; Haeuber, R: Huber, CM: Lynch,
L (2012b). Critical loads of acidity to protect and restore acid-sensitive streams in
Virginia and West Virginia. Water Air Soil Pollut 223: 5759-5771.
http://dx.doi.org/10.1007/sll270-Q12-1312-4.
Sullivan, TJ: Cosby, BJ: Tonnessen, KA: Clow, DW. (2005). Surface water acidification
responses and critical loads of sulfur and nitrogen deposition in Loch Vale Watershed,
Colorado. Water Resour Res 41: W01021. http://dx.doi.org/10.1029/2004WR0Q3414.
Sullivan, TJ: Cosby, BJ: Webb, JR: Dennis, RL: Bulger, AJ: Devinev FA, J, r. (2008).
Streamwater acid-base chemistry and critical loads of atmospheric sulfur deposition in
Shenandoah National Park, Virginia. Environ Monit Assess 137: 85-99.
http://dx.doi.org/10.1007/slQ661-007-9731-l.
Sullivan, TJ: Cosby, BJ: Webb, JR: Snyder, KU: Herlihy, AT: Bulger, AJ: Gilbert, EH: Moore,
D. (2002). Assessment of the effects of acidic deposition on aquatic resources in the
Southern Appalachian Mountains. Corvallis, OR: E&S Environmental Chemistry, Inc.
http://people.virginia.edu/~swas/POST/assets/docs/Acid Pep Effects aquatic final rep
ort 0702.pdf.
Sullivan, TJ: Driscoll, CT: Cosby, BJ: Fernandez, IJ: Herlihy, AT: Zhai, J: Stemberger, R:
Snyder, KU: Sutherland, JW: Nierzwicki-Bauer, SA: Boylen, CW: McDonnell, TC:
Nowicki, NA. (2006a). Assessment of the extent to which intensively-studied lakes are
representative of the Adirondack Mountain region. Final report. Corvallis, OR: E&S
Environmental Chemistry, Inc.
Sullivan, TJ: Fernandez, IJ: Herlihy, AT: Driscoll, CT: McDonnell, TC: Nowicki, NA: Snyder,
KU: Sutherland, JW. (2006b). Acid-base characteristics of soils in the Adirondack
Mountains, New York. Soil Sci Soc Am J 70: 141-152.
http://dx.doi.org/10.2136/sssai2005.0Q73.
Sullivan, TJ: Jenkins, J. (2014). The science and policy of critical loads of pollutant deposition to
protect ecosystems in New York. Ann NY Acad Sci 1313: 57-68.
http://dx.doi.org/10. Ill 1/nyas. 12407.
June 2018
R-205
DRAFT: Do Not Cite or Quote

-------
Sullivan. TJ; Lawrence. GB; Bailey. SW; McDonnell TC; Beier, CM; Weathers. KC;
McPherson. GT; Bishop. DA. (2013). Effects of acidic deposition and soil acidification
on sugar maple trees in the Adirondack Mountains, New York. Environ Sci Technol 47:
12687-12694. http://dx.doi.org/10.1021/es4Q1864w.
Sullivan. TJ: Webb. JR; Snyder. KU; Herlihy. AT: Cosby. BJ. (2007b). Spatial distribution of
acid-sensitive and acid-impacted streams in relation to watershed features in the southern
Appalachian mountains. Water Air Soil Pollut 182: 57-71.
http://dx.doi.org/10.1007/sll270-006-932Q-x.
Summers. JC: Kurek. J: Kirk. JL; Muir. DC: Wang. X: Wiklund. JA; Cooke. CA; Evans. MS:
Smol. JP. (2016). Recent Warming, Rather than Industrial Emissions of Bioavailable
Nutrients, Is the Dominant Driver of Lake Primary Production Shifts across the
Athabasca Oil Sands Region. PLoS ONE 11: e0153987.
http://dx.doi.org/10.1371/iournal.pone.0153987.
Sun. K; Cadv-Pereira. K; Miller. DJ; Tao. L; Zondlo. MA; Nowak. JB; Neuman. JA; Mikoviny.
T; Muller, M; Wisthaler, A; Scarino, AJ; Hostetler, CA. (2015a). Validation of TES
ammonia observations at the single pixel scale in the San Joaquin Valley during
DISCOVER-AQ. J Geophys Res Atmos 120: 5140-5154.
http://dx.doi.org/10.1002/2014JDQ22846.
Sun, L; Lin, Z; Liao, K; Xi, Z; Wang, D. (2015b). Adverse effects of coal combustion related
fine particulate matter (PM2.5) on nematode Caenorhabditis elegans. Sci Total Environ
512-513: 251-260. http://dx.doi.Org/10.1016/i.scitotenv.2015.01.058.
Sun, L; Wu, O; Liao, K, ai; Yu, P; Cui, O; Rui, O, i; Wang, D. (2016). Contribution of heavy
metals to toxicity of coal combustion related fine particulate matter (PM2.5) in
Caenorhabditis elegans with wild-type or susceptible genetic background. Chemosphere
144: 2392-2400. http://dx.doi.org/10.1016/i.chemosphere.2015.11.028.
Sun, S; Xing, F; Zhao, H; Gao, Y; Bai, Z; Dong, Y. (2014). Response of bacterial community to
simulated nitrogen deposition in soils and a unique relationship between plant species and
soil bacteria in the Songnen grassland in Northeastern China. JSSPN 14: 565-580.
http://dx.doi.org/10.4067/S0718-95162014005000Q45.
Sunda, WG; Cai, W. (2012). Eutrophication induced C02-acidification of subsurface coastal
waters: Interactive effects of temperature, salinity, and atmospheric P-C02. Environ Sci
Technol 46: 10651-10659. http://dx.doi.org/10.1021/es300626f.
Sundberg, B; Nasholm, T; Palmqvist, K. (2001). The effect of nitrogen on growth and key
thallus components in the two tripartite lichens, Nephroma arcticum and Peltigera
aphthosa. Plant Cell Environ 24: 517-527. http://dx.doi.org/10.1046/j. 1365-
3040.2001.00701.x.
Sundin, J; Aronsen, T; Rosenqvist, G; Berglund, A. (2017). Sex in murky waters: algal-induced
turbidity increases sexual selection in pipefish. Behav Ecol Sociobiol 71: 78.
http://dx.doi.org/10.1007/sQ0265-017-2310-8.
Sundin, J; Berglund, A; Rosenqvist, G. (2010). Turbidity hampers mate choice in a pipefish.
Ethology 116: 713-721. http://dx.doi.org/10.1111/i. 1439-0310.2010.01787.x.
June 2018
R-206
DRAFT: Do Not Cite or Quote

-------
Sundin. J; Rosenqvist G; Berglund. A. (2015). Hypoxia delays mating in the broad-nosed
pipefish. Mar Biol Res 11: 747-754. http://dx.doi.org/10.1080/17451000.2015.10Q7874.
Sundqvist MK; Liu. Z; Giesler. R; Wardle. DA. (2014). Plant and microbial responses to
nitrogen and phosphorus addition across an elevational gradient in subarctic tundra.
Ecology 95: 1819-1835. http://dx.doi.Org/10.1890/13-0869.l.
Suter. GW: Norton. SB: Fairbrother. A. (2005). Individuals versus organisms versus populations
in the definition of ecological assessment endpoints. Integr Environ Assess Manag 1:
397-400. http://dx.doi.org/10.1002/ieam.56300104Q9.
Sutherland, JW; Acker, FW; Bloomfield, JA; Boylen, CW; Charles, DF; Daniels, RA; Eichler,
LW: Farrell, JL; Feranec, RS: Hare, MP: Kanfoush, SL; Preall, RJ: Quinn, SO: Rowell,
HC: Schoch, WF: Shaw, WH: Siegfried, CA: Sullivan, TJ: Winkler, DA: Nierzwicki-
Bauer, SA. (2015). Brooktrout lake case study: Biotic recovery from acid deposition 20
years after the 1990 clean air act amendments. Environ Sci Technol 49: 2665-2674.
http://dx.doi.org/10.1021/es5036865.
Sutton, MA: Burkhardt, JK; Guerin, D; Nemitz, E; Fowler, D. (1998). Development of resistance
models to describe measurements of bi-directional ammonia surface-atmosphere
exchange. Atmos Environ 32: 473-480. http://dx.doi.Org/10.1016/S 1352-2310(97)00164-
7.
Sutton. MA: Fowler. D: Moncrieff, JB. (1993). THE EXCHANGE OF ATMOSPHERIC
AMMONIA WITH VEGETATED SURFACES .1. UNFERTILIZED VEGETATION. Q
J Roy Meteorol Soc 119: 1023-1045.
Sutton. MA: Howard. CM: Erisman. JW: Billen. G: Bleeker. A: Grennfelt. P: van Grinsven. H:
Grizzetti, B. (2011). The European nitrogen assessment: Sources, effects and policy
perspectives. In MA Sutton; CM Howard; JW Erisman; G Billen; A Bleeker; P Grennfelt;
H van Grinsven; B Grizzetti (Eds.), The European Nitrogen Assessment: Sources, Effects
and Policy Perspectives (pp. 664). Cambridge, UK: Cambridge University Press.
Sutton. MA: Schiorring. JK: Wvers. GP. (1995). PLANT ATMOSPHERE EXCHANGE OF
AMMONIA. Philos Transact A Math Phys Eng Sci 351: 261-276.
Sutton. MA: Simpson. D: Levy. PE: Smith. RI; Reis. S: van Oiien. M: de Vries. W. (2008).
Uncertainties in the relationship between atmospheric nitrogen deposition and forest
carbon sequestration. Global Change Biol 14: 2057 - 2063.
http://dx.doi.org/10.1111/i. 1365-2486.2008.01636.x.
Suz. LM: Barsoum. N: Benham. S: Dietrich. HP: Fetzer. KD: Fischer. R: Garcia. P: Gehrman. J:
Kristofel. F: Manninger. M: Neagu. S: Nicolas. M: Oldenburger. J: Raspe. S: Sanchez. G:
Schrock. HW: Schubert. A: Verheven. K: Verstraeten. A: Bidartondo. MI. (2014).
Environmental drivers of ectomycorrhizal communities in Europe's temperate oak
forests. Mol Ecol 23: 5628-5644. http://dx.doi.org/10. Ill 1/mec. 12947.
Svendsen. ML: Steinnes. E: Blom. HA. (2011). Partitioning of Zn, Cd, Pb, and Cu in organic-
rich soil profiles in the vicinity of a zinc smelter. Chem Speciation Bioavailability 23:
189-200. http://dx.doi.org/10.3184/095422911X13103862613Q85.
June 2018
R-207
DRAFT: Do Not Cite or Quote

-------
Sverdrup. H; Belvazid. S; Nihlgard. B; Ericson. L. (2007). Modelling change in ground
vegetation response to acid and nitrogen pollution, climate change and forest
management at in Sweden 15002100 A.D. Water, Air, and Soil Pollution: Focus 7: 163-
179. http://dx.doi.org/10.1007/sll267-006-9Q67-9.
Sverdrup. H; McDonnell. TC: Sullivan. TJ; Nihlgard. B; Belvazid. S: Rihm. B; Porter. E;
Bowman. WD: Geiser. L. (2012). Testing the feasibility of using the ForSAFE-VEG
model to map the critical load of nitrogen to protect plant biodiversity in the Rocky
Mountains Region, USA. Water Air Soil Pollut 223: 371-387.
http://dx.doi.org/10.1007/sll270-011-Q865-v.
Sverdrup. H; Warfvinge. P. (1993). The effect of soil acidification on the growth of trees, grass
and herbs as expressed by the (Ca+ Mg+ K)/A1 ratio (2nd ed.). Lund, Sweden: KF-
Sigma.
Svobodova. J: Matena. J: Kopacek. J: Polakova. S: Vrba. J. (2012). Spatial and temporal changes
of benthic macroinvertebrate assemblages in acidified streams in the Bohemian Forest
(Czech Republic). Aquatic Insects 34: 157-172.
http://dx.doi.org/10.1080/01650424.2012.643Q48.
Swain. RE. (1949). Smoke and fume investigations: A historical review. In Atmospheric
Contamination and Purification. Washington, DC: American Chemical Society.
http://dx.doi.org/10.1021/ie50479a012.
Swanev. DP: Scavia. D; Howarth. RW: Marino. RM. (2008). Estuarine classification and
response to nitrogen loading: Insights from simple ecological models. Estuar Coast Shelf
Sci 77: 253-263. http://dx.doi.Org/10.1016/i.ecss.2007.09.013.
Symons. CC: Arnott. SE; Sweetman. J. (2012). Nutrient limitation of phytoplankton
communities in Subarctic lakes and ponds in Wapusk National Park, Canada. Polar
Biology 35: 481-489. http://dx.doi.org/10.1007/s00300-011-1092-Q.
Svphard. AD: Keelev. J: Abatzoglou. JT. (2017). Trends and drivers of fire activity vary across
California aridland ecosystems. J Arid Environ 104.
http://dx.doi.Org/10.1016/i.iaridenv.2017.03.017.
Taboada, A; Marcos, E; Calvo, L. (2016). Disruption of trophic interactions involving the
heather beetle by atmospheric nitrogen deposition. Environ Pollut 218: 436-445.
http://dx.doi.Org/10.1016/i.envpol.2016.07.023.
Tagaris. E; Liao. KJ; Manomaiphiboon. K; Woo. JH; He. S: Amar. P; Russell. AG. (2008).
Impacts of future climate change and emissions reductions on nitrogen and sulfur
deposition over the United States. Geophys Res Lett 35: L08811.
http://dx.doi.org/10.1029/2008GLQ33477.
Talbot. JM; Treseder. KK. (2012). Interactions among lignin, cellulose, and nitrogen drive litter
chemistry-decay relationships. Ecology 93: 345-354. http://dx.doi.org/10.189Q/ll-
0843.1.
Talbot. RW: Vijgen. AS: Harriss. RC. (1990). Measuring tropospheric HN03: Problems and
prospects for nylon filter and mist chamber techniques. J Geophys Res 95: 7553-7561.
http://dx.doi.org/10.1029/JD095iD06pQ7553.
June 2018
R-208
DRAFT: Do Not Cite or Quote

-------
Talhelm. AF; Burton. AJ; Pregitzer. KS; Campione. MA. (2013). Chronic nitrogen deposition
reduces the abundance of dominant forest understory and groundcover species. For Ecol
Manage 293: 39-48. http://dx.doi.Org/10.1016/i.foreco.2012.12.020.
Talhelm. AF: Pregitzer. KS: Burton. AJ. (2011). No evidence that chronic nitrogen additions
increase photosynthesis in mature sugar maple forests. Ecol Appl 21: 2413-2424.
http://dx.doi.Org/10.1890/10-2076.l.
Talhelm. AF: Pregitzer. KS: Burton. AJ: Zak. DR. (2012). Air pollution and the changing
biogeochemistry of northern forests. Front Ecol Environ 10: 181-185.
http://dx.doi.Org/10.1890/l 10007.
Talhelm. AF: Pregitzer. KS: Zak. DR. (2009). Species-specific responses to atmospheric carbon
dioxide and tropospheric ozone mediate changes in soil carbon. Ecol Lett 12: 1219-1228.
http://dx.doi.org/10.111 l/i.l461-0248.2009.01380.x.
Talluto, MY: Suding, KN. (2008). Historical change in coastal sage scrub in southern California,
USA in relation to fire frequency and air pollution. Landsc Ecol 23: 803-815.
http://dx.doi.org/10.1007/slQ980-008-9238-3.
Tang. W: Cohan. DS: Pour-Biazar. A: Lamsal. LN: White. AT: Xiao. X: Zhou. W: Henderson.
BH: Lash, BF. (2015). Influence of satellite-derived photolysis rates and NOx emissions
on Texas ozone modeling. Atmos Chem Phys 15: 1601-1619.
http://dx.doi.org/10.5194/acp-15-1601-2015.
Taulavuori. K: Prasad. MNV: Taulavuori. E: Laine. K. (2005). Metal stress consequences on
frost hardiness of plants at northern high latitudes: a review and hypothesis [Review],
Environ Pollut 135: 209-220. http://dx.doi.org/10.1016/i.envpol.2004.11.006.
Taylor. OC: MacLean. DC. (1970). Nitrogen oxides and the peroxyacyl nitrates. In JS Jacobson;
AC Hill (Eds.), Recognition of Air Pollution Injury to Vegetation: A Pictorial Atlas (pp.
E1-E14). Pittsburgh, PA: Air Pollution Control Association.
TBEP (Tampa Bay Estuary Program). (2006). Charting the course: The comprehensive
conservation and management plan for Tampa Bay. St. Petersburg, FL.
TBEP (Tampa Bay Estuary Program). (2011). 2010 Tampa Bay Water quality assessment.
(Technical Report #01-11). St. Petersburg, FL.
http://www.tbeptech.org/TBEP TECH PUBS/2011/TBEP 01 11 2010 Decision Matri
x Results Update.pdf.
TBEP (Tampa Bay Estuary Program). (2012a). Tampa Bay nitrogen management strategy 2012
reasonable assurance update document. St. Petersburg, FL.
http://www.tbeptech.org/NitrogenMgmtConsort/ReasonableAssurance/2012 Submittal/F
INAL 2012 RA Update NMC Approved DEC142012.pdf.
TBEP (Tampa Bay Estuary Program). (2012b). Understanding atmospheric deposition in Tampa
Bay. Available online at
http://www.tbeptech.org/TBEP TECH PUBS/2012/TBEP 04 12 BRACE Public Doc
ument.pdf (accessed
June 2018
R-209
DRAFT: Do Not Cite or Quote

-------
TBEP (Tampa Bay Estuary Program). (2014). Tampa bay estuary program: 2014 2015 EPA
workplan and budget. St. Petersburg, FL.
http://tbep.org/pdfs/Full FINAL April 25 2014.pdf.
TBEP (Tampa Bay Estuary Program). (2015). A Tampa Bay Estuary Program progress report.
Available online at http://www.tbep.org/pdfs/tbep progress rpt 2015 webx.pdf
(accessed
Temple. PJ: Taylor. OC. (1983). World-wide ambient measurements of peroxyacetyl nitrate
(PAN) and implications for plant injury. Atmos Environ 17: 1583-1587.
http://dx.doi.org/10.1016/0004-6981(83^)90311-6.
Templer. PH: Lovett. GM; Weathers. KC: Findlav. SE; Dawson. TE. (2005). Influence of tree
species on forest nitrogen retention in the Catskill Mountains, New York, USA.
Ecosystems 8: 1-16. http://dx.doi.org/10.1007/sl0021-004-0230-8.
Templer, PH; Mack, MC; Chapin, FS, III; Christenson, LM; Compton, JE; Crook, HP; Currie,
WS; Curtis, CJ; Pail, DB; D'Antonio, CM; Emmett, BA; Epstein, HE; Goodale, CL;
Gundersen, P; Hobbie, SE; Holland, K; Hooper, DU; Hungate, BA; Lamontagne, S;
Nadelhoffer, KJ; Osenberg, CW; Perakis, SS; Schleppi, P; Schimel, J; Schmidt, IK;
Sommerkorn, M; Spoelstra, J; Tietema, A; Wessel, WW; Zak, DR. (2012). Sinks for
nitrogen inputs in terrestrial ecosystems: A meta-analysis of N-15 tracer field studies.
Ecology 93: 1816-1829. http://dx.doi.Org/10.1890/ll-1146.l.
Templer, PH; Weathers, KC; Ewing, HA; Dawson, TE; Mambelli, S; Lindsev, AM; Webb, J;
Boukili, VK; Firestone, MK. (2015a). Fog as a source of nitrogen for redwood trees:
Evidence from fluxes and stable isotopes. J Ecol 103: 1397-1407.
http://dx.doi.org/10.1111/1365-2745.12462.
Templer, PH; Weathers, KC; Lindsev, A; Lenoir, K; Scott, L. (2015b). Atmospheric inputs and
nitrogen saturation status in and adjacent to Class I wilderness areas of the northeastern
US. Oecologia 177: 5-15. http://dx.doi.org/10.1007/s00442-014-3121-5.
Teskev, RO; Gholz, HL; Cropper, WP, Jr. (1994). Influence of climate and fertilization on net
photosynthesis of mature slash pine. Tree Physiol 14: 1215-1227.
http://dx.doi.org/10.1093/treephys/14.ll.1215.
Tessler, M; Truhn, KM; Bliss-Moreau, M; Wehr, JD. (2014). Diversity and distribution of
stream bryophytes: Does pH matter? Freshw Sci 33: 778-787.
http://dx.doi.org/10.1086/676996.
Testa, JM; Clark, JB; Dennison, WC; Donovan, EC; Fisher, AW; Ni, W; Parker, M; Scavia, D;
Spitzer, SE; Waldrop, AM; Vargas, VMD; Ziegler, G. (2017). Ecological Forecasting
and the Science of Hypoxia in Chesapeake Bay. Bioscience 67: 614-626.
http://dx.doi.org/10.1093/biosci/bix048.
Testa, JM; Kemp, WM. (2012). Hypoxia-induced shifts in nitrogen and phosphorus cycling in
Chesapeake Bay. Limnol Oceanogr 57: 835-850.
http://dx.doi.Org/10.4319/lo.2012.57.3.0835.
June 2018
R-210
DRAFT: Do Not Cite or Quote

-------
Thevs. N; De Smedt I; van Gent J: Danckaert T; Wang. T; Hendrick. F; Stavrakou. T; Bauduin.
S; Clarisse. L; Li. C; Krotkov. N; Yu. H; Brenot H; Van Roozendael, M. (2015). Sulfur
dioxide vertical column DOAS retrievals from the Ozone Monitoring Instrument: Global
observations and comparison to ground-based and satellite data. J Geophys Res Atmos
120: 2470-2491. http://dx.doi.org/10.1002/2014JDQ22657.
Thibodeau. B; Helie. JF; Lehmann. MF. (2013). Variations of the nitrate isotopic composition in
the St. Lawrence River caused by seasonal changes in atmospheric nitrogen inputs.
Biogeochemistry 115: 287-298. http://dx.doi.org/10.1007/slQ533-013-9834-4.
Thimonier. A: Pannatier. EG: Schmitt. M; Waldner. P; Walthert. L; Schleppi. P; Dobbertin. M;
Krauchi. N. (2010). Does exceeding the critical loads for nitrogen alter nitrate leaching,
the nutrient status of trees and their crown condition at Swiss Long-term Forest
Ecosystem Research (LWF) sites? European Journal of Forest Research 129: 443-461.
http://dx.doi.org/10.1007/slQ342-009-0328-9.
Thoene. B; Schroder. P; Papen. H: Egger. A: Rennenberg. H. (1991). Absorption of atmospheric
N02 by spruce (Picea abies L. Karst.) trees. I. N02 influx and its correlation with nitrate
reduction. NewPhytol 117: 575-585. http://dx.doi.Org/10.l 11 l/j.1469-
8137.199 l.tb00962.x.
Thomas. GO: Smith. KEC: Sweetman. AJ; Jones. KC. (1998). Further studies of the air-pasture
transfer of polychlorinated biphenyls. Environ Pollut 102: 119128.
http://dx.d0i.0rg/l 0.1016/S0269-7491 (98)00070-0.
Thomas. MP. (1951). Gas damage to plants. Annu Rev Plant Biol 2: 293-322.
http://dx.doi.Org/10.l 146/annurev. pp. 02.060151.001453.
Thomas. MP: Hendricks. RH: Hill. GR. (1944). Some chemical reactions of sulfur dioxide after
absorption by alfalfa and sugarbeets. J Plant Physiol 19: 212-226.
http://dx.doi.Org/10.1104/pp.19.2.212.
Thomas. P; Rahman. MS. (2009). Biomarkers of hypoxia exposure and reproductive function in
Atlantic croaker: A review with some preliminary findings from the northern Gulf of
Mexico hypoxic zone. Exp Mar Bio Ecol 381: S38-S50.
http://dx.doi.Org/10.1016/i.iembe.2009.07.008.
Thomas. P; Rahman. MS: Khan. IA; Kummer. JA. (2007). Widespread endocrine disruption and
reproductive impairment in an estuarine fish population exposed to seasonal hypoxia.
Proc Biol Sci 274: 2693-2701. http://dx.doi.org/10.1098/rspb.2007.Q921.
Thomas, RB; Spal, SE; Smith, KR; Nippert, JB. (2013). Evidence of recovery of Juniperus
virginiana trees from sulfur pollution after the Clean Air Act. Proc Natl Acad Sci USA
110: 15319-15324. http://dx.doi.org/10.1073/pnas.1308115110.
Thomas, RO; Brookshire, EN: Gerber, S. (2015). Nitrogen limitation on land: How can it occur
in Earth system models? [Review], Global Change Biol 21: 1777-1793.
http://dx.doi.org/10. Ill 1/gcb. 12813.
Thomas, RO: Canham, CP: Weathers, KC: Goodale, CL. (2010). Increased tree carbon storage
in response to nitrogen deposition in the US. Nat Geosci 3: 13-17.
http://dx.doi.org/10.1038/NGEQ721.
June 2018
R-211
DRAFT: Do Not Cite or Quote

-------
Thompson. CR; Hensel. EG; Kats. G; Taylor. OC. (1970). Effects of continuous exposure of
navel oranges to nitrogen dioxide. Atmos Environ 4: 349-355.
http://dx.d0i.0rg/l 0.1016/0004-6981 (70)90079-X.
Thompson. PL; St-Jacques. M. -C; Vinebrooke. RD. (2008). Impacts of climate warming and
nitrogen deposition on alpine plankton in lake and pond habitats: An in vitro experiment.
Arctic Antarct Alpine Res 40: 192-198. http://dx.doi.org/10.1657/1523-0430(06-
105)rTHOMPSON12.Q.CO;2.
Thompson. TM; Rodriguez. MA; Barna. MG; Gebhart. KA; Hand. JL; Day. DE; Malm. WC;
Benedict. KB; Collett. JL. Jr; Schichtel. BA. (2015). Rocky Mountain National Park
reduced nitrogen source apportionment. J Geophys Res Atmos 120: 4370-4384.
http://dx.doi. org/10.1002/2014JD022675.
Thornberry-Ehrlich. T. (2008). Great Smoky Mountains National Park : Geologic resource
evaluation report. (Natural Resource Report NPS/NRPC/GRD/NRR2008/048).
Washington, D.C.: U.S. Department of the Interior, Geologic Resources Division,
Natural Resource Program Center.
Thornton. PE; Donev. SC; Lindsay. K; Moore. JK; Mahowald. N; Randerson. JT; Fung. I;
Lamarque. J. -F; Feddema. JJ; Lee. Y. -H. (2009). Carbon-nitrogen interactions regulate
climate-carbon cycle feedbacks: Results from an atmosphere-ocean general circulation
model. Biogeosciences 6: 2099-2120. http://dx.doi.org/10.5194/bg-6-2099-2009.
Thornton. PE; Lamarque. J. -F; Rosenbloom. NA; Mahowald. NM. (2007). Influence of carbon-
nitrogen cycle coupling on land model response to CO(2) fertilization and climate
variability. Global Biogeochem Cycles 21: GB4018.
http://dx.doi.org/10.1029/2006GB0Q2868.
Throop. HL; Lerdau. MT. (2004). Effects of nitrogen deposition on insect herbivory:
Implications for community and ecosystem processes. Ecosystems 7: 109-133.
http://dx.doi. org/10.1007/s 10021 -003 -0225-x.
Tian. D; Wang. H; Sun. J; Niu. S. (2016a). Global evidence on nitrogen saturation of terrestrial
ecosystem net primary productivity. Environ Res Lett 11. http://dx.doi.org/10.lQ88/1748-
9326/11/2/024012.
Tian. H; Chen. G; Zhang. C; Liu. M; Sun. G; Chappelka. A; Ren. W; Xu. X; Lu. C; Pan. S;
Chen. H; Hui. D; McNultv. S; Lockabv. G; Vance. E. (2012). Century-scale responses of
ecosystem carbon storage and flux to multiple environmental changes in the southern
United States. Ecosystems 15: 674-694. http://dx.doi.org/10.1007/slQ021-012-9539-x.
Tian. H; Melillo. J; Lu. C; Kicklighter. D; Liu. M; Ren. W; Xu. X; Chen. G; Zhang. C; Pan. S;
Liu. J; Running. S. (2011). China's terrestrial carbon balance: Contributions from
multiple global change factors. Global Biogeochem Cycles 25: GB1007.
http://dx.doi.org/10.1029/2010GB0Q3838.
Tian. Q; Liu. N; Bai. W; Li. L; Chen. J; Reich. PB; Yu. Q; Guo. D; Smith. MP; Knapp. AK;
Cheng. W; Lu. P; Gao. Y; Yang. A; Wang. T; Li. X; Wang. Z; Ma. Y; Han. X; Zhang.
W. (2016b). A novel soil manganese mechanism drives plant species loss with increased
nitrogen deposition in a temperate steppe. Ecology 97: 65-74.
http ://dx. doi. org/10.1890/15-0917.1.
June 2018
R-212
DRAFT: Do Not Cite or Quote

-------
Tietema. A; Emmett BA; Gundersen. P; Ki0naas, OJ; Koopmans. CJ. (1998). The fate of 15N-
labelled nitrogen deposition in coniferous forest ecosystems. For Ecol Manage 101: 19-
27. http://dx.doi.org/10.1016/SQ378-l 127(97)00123-0.
Tilman, D. (1987). Secondary succession and the pattern of plant dominance along experimental
nitrogen gradients. Ecol Monogr 57: 189-214. http://dx.doi.org/10.2307/2937080.
Tilman, D. (1993). Species richness of experimental productivity gradients: How important is
colonization limitation? Ecology 74: 2179-2191. http://dx.doi.org/10.2307/1939572.
Tilman, D. (1996). Biodiversity: Population versus ecosystem stability. Ecology 77: 350-363.
http://dx.doi.org/10.2307/2265614.
Tilman, D. (2000). Causes, consequences and ethics of biodiversity. Nature 405: 208-211.
http://dx.doi.org/10.1038/35012217.
Tilman, D; Downing. JA. (1994). Biodiversity and stability in grasslands. Nature 367: 363-365.
http://dx.doi.org/10.1038/367363a0.
Tilman, D; Knops, J: Wedin, D; Reich. P; Ritchie. M; Siemann, E. (1997). The influence of
functional diversity and composition on ecosystem processes. Science 277: 1300-1302.
http://dx.doi.Org/10.l 126/science.277.5330.1300.
Tilman. D; Lehman. CL; Bristow. CE. (1998). Diversity-stability relationships: Statistical
inevitability or ecological consequence? Am Nat 151: 277-282.
Tilman. D; Reich. PB; Isbell. F. (2012). Biodiversity impacts ecosystem productivity as much as
resources, disturbance, or herbivory. Proc Natl Acad Sci USA 109: 10394-10397.
http://dx.doi.org/10.1073/pnas. 1208240109.
Tilman. D; Reich. PB: Knops. J: Wedin. D; Mielke. T; Lehman. C. (2001). Diversity and
productivity in a long-term grassland experiment. Science 294: 843-845.
http ://dx. doi .org/10.1126/science. 10603 91.
Tilman, D; Reich, PB; Knops, JMH. (2006). Biodiversity and ecosystem stability in a decade-
long grassland experiment. Nature 441: 629-632. http://dx.doi.org/10.1038/nature04742.
Tilman, D; Wedin, D. (1991). Dynamics of nitrogen competition between successional grasses.
Ecology 72: 1038-1049. http://dx.doi.org/10.2307/1940604.
Tipping, E; Benham, S: Boyle, JF; Crow, P; Davies, J: Fischer, U; Guvatt, Ft; Helliwell, R;
Jackson-Blake, L; Lawlor, AJ; Monteith, DT; Rowe, EC: Toberman, H. (2014).
Atmospheric deposition of phosphorus to land and freshwater. Environ Sci Process
Impacts 16: 1608-1617. http://dx.doi.org/10.1039/c3em00641g.
Tipping, E; Davies, JAC: Henrys, PA: Kirk, GJD; Lilly, A: Dragosits, U: Carnell, EJ; Pore, AJ:
Sutton, MA: Tomlinson, SJ. (2017). Long-term increases in soil carbon due to ecosystem
fertilization by atmospheric nitrogen deposition demonstrated by regional-scale
modelling and observations. Sci Rep 7: 1890. http://dx.doi.org/10.1038/s41598-Q17-
02002-w.
June 2018
R-213
DRAFT: Do Not Cite or Quote

-------
Tipping. E; Rowe. EC; Evans. CD; Mills. RTE; Emmett BA; Chaplow. JS; Hall. JR. (2012).
N14C: A plant-soil nitrogen and carbon cycling model to simulate terrestrial ecosystem
responses to atmospheric nitrogen deposition. Ecol Modell 247: 11-26.
http://dx.doi.Org/10.1016/i.ecolmodel.2012.08.002.
Titus. JH; Nowak. RS; Smith. SD. (2002). Soil resource heterogeneity in the Mojave Desert. J
Arid Environ 52: 269-292. http://dx.doi.org/10.1006/iare.2002.101Q.
Tierngren. I; Karlsson. T; Biorn. E; Skyllberg. U. (2012). Potential Hg methylation and MeHg
demethylation rates related to the nutrient status of different boreal wetlands.
Biogeochemistry 108: 335-350. http://dx.doi.org/10.1007/slQ533-011-9603-1.
Todorova. SG; Driscoll. CT; Effler. SW; O'Donnell. S; Matthews. DA; Todorov. PL;
Gindlesperger. S. (2014). Changes in the long-term supply of mercury species to the
upper mixed waters of a recovering lake. Environ Pollut 185: 314-321.
http://dx.doi.Org/10.1016/i.envpol.2013.l 1.005.
Todorova. SG; Driscoll. CT; Matthews. DA; Effler. SW; Hines. ME; Henry. EA. (2009).
Evidence for regulation of monomethyl mercury by nitrate in a seasonally stratified,
eutrophic lake. Environ Sci Technol 43: 6572-6578. http://dx.doi.org/10.1021/es900887b.
Tolocka, MP; Turpin, B. (2012). Contribution of organosulfur compounds to organic aerosol
mass. Environ Sci Technol 46: 7978-7983. http://dx.doi.org/10.1021/es300651v.
Tomasko. DA; Corbett. CA; Greening. HS; Raulerson. GE. (2005). Spatial and temporal
variation in seagrass coverage in Southwest Florida: Assessing the relative effects of
anthropogenic nutrient load reductions and rainfall in four contiguous estuaries. Mar
Pollut Bull 50: 797-805. http://dx.doi.Org/10.1016/i.marpolbul.2005.02.010.
Tomassen. HBM; Smolders. AJP; Lamers. LPM; Roelofs. JGM. (2003). Stimulated growth of
Betula pubescens and Molinia caerulea on ombrotrophic bogs: Role of high levels of
atmospheric nitrogen deposition. J Ecol 91: 357-370. http://dx.doi.org/10.1046/j. 1365-
2745.2003.00771.x.
Tominaga. K; Aherne. J; Watmough. SA; Alveteg. M; Cosby. BJ; Driscoll. CT; Posch. M.
(2009). Voyage without constellation: Evaluating the performance of three uncalibrated
process-oriented models. 40: 261-272. http://dx.doi.org/10.2166/nh.2009.085.
Tominaga. K; Aherne. J; Watmough. SA; Alveteg. M; Cosby. BJ; Driscoll. CT; Posch. M;
Pourmokhtarian. A. (2010). Predicting acidification recovery at the Hubbard Brook
Experimental Forest, New Hampshire: Evaluation of four models. Environ Sci Technol
44: 9003-9009. http://dx.doi.org/10.1021/esl02243i.
Tong. C; She. CX; Yang. P; Jin. YF; Huang. JF. (2015). Weak correlation between methane
production and abundance of methanogens across three brackish marsh zones in the Min
River Estuary, China. Estuaries Coasts 38: 1872-1884. http://dx.doi.org/10.10Q7/sl2237-
014-9930-2.
Tonitto. C; Goodale. CL; Weiss. MS; Frev. SD; Ollinger. SV. (2014). The effect of nitrogen
addition on soil organic matter dynamics: A model analysis of the Harvard Forest
Chronic Nitrogen Amendment Study and soil carbon response to anthropogenic N
deposition. Biogeochemistry 117: 431-454. http://dx.doi.org/10.1007/slQ533-013-9887-
4.
June 2018
R-214
DRAFT: Do Not Cite or Quote

-------
Tonnesen, G; Wang. Z; Omary, M; Chien, CJ. (2007). Assessment of nitrogen deposition:
Modeling and habitat assessment. (CEC-500-2005-032). Sacramento, CA: California
Energy Commission, PIER Energy-Related Environmental Research.
http://www.energv.ca.gov/2006publications/CEC-500-2006-032/CEC-500-20Q6-
032.PDF.
Townhill. BL; Pinnegar. JK: Righton. DA: Metcalfe. JD. (2017). Fisheries, low oxygen and
climate change: how much do we really know? [Review], J Fish Biol 90: 723-750.
http://dx.doi.org/10. Ill 1/ifb. 13203.
Townsend. AR; Howarth. RW: Bazzaz. FA: Booth. MS: Cleveland. CC: Collinge. SK; Dobson.
AP; Epstein. PR: Holland. EA: Keenev. PR: Mallin. MA: Rogers. CA: Wayne. P; Wolfe.
AH. (2003). Human health effects of a changing global nitrogen cycle. Front Ecol
Environ 1: 240-246. http://dx.doi.org/10.1890/154Q-
9295(2003X)0ir0240:HHEOAC12.0.CO:2.
Trainer. VL: Bates. SS: Lundholm. N: Thessen. AE: Cochlan. WP: Adams. NG: Trick. CG.
(2012). Pseudo-nitzschia physiological ecology, phylogeny, toxicity, monitoring and
impacts on ecosystem health. Harmful Algae 14: 271-300.
http://dx.doi.Org/10.1016/i.hal.2011.10.025.
Traister. EM: McDowell. WH: Kram. P; Fottova. D: Kolarikova. K. (2013). Persistent effects of
acidification on stream ecosystem structure and function. Freshw Sci 32: 586-596.
http://dx.doi.Org/10.1899/12-130.l.
Travis. KR. Jacob. D. J.: Fisher. JA: Kim. PS: Marais. EA: Zhu. L: Yu. K: Miller. CC: Yantosca.
RM: Sulprizio. MP: Thompson. AM: Wennberg. PO: Crounse. JD: St. Clair. JM: Cohen.
RC: Laughner, JL: Dibb, JE: Hall, SR: Ullmann, K: Wolfe, GM: Neuman, JA: Zhou, X.
(2016).	Why do models overestimate surface ozone in the Southeastern United States?
Atmos Chem Phys Discuss 16: 13561-13577. http://dx.doi.org/10.5194/acp-2016-l 10.
Treseder, KK. (2004). The meta-analysis of mycorrhizal response to nitrogen, phosphorus, and
atmospheric C02 in field studies. NewPhytol 164: 347-355.
http://dx.doi.org/10.1111/i. 1469-8137.2004.01159.x.
Treseder, KK. (2008). Nitrogen additions and microbial biomass: A meta-analysis of ecosystem
studies. Ecol Lett 11: 1111-1120. http://dx.doi.Org/10.l 111/i. 1461-0248.2008.01230.x.
Tritton, LM: Martin, CW: Hornbeck, JW: Pierce, RS. (1987). Biomass and nutrient removals
from commercial thinning and whole-tree clearcutting of central hardwoods. Environ
Manage 11: 659-666. http://dx.doi.org/10.1007/BF0188Q165.
Trommer, G: Levnaert, A: Klein, C: Naegelen, A: Beker, B. (2013). Phytoplankton phosphorus
limitation in a North Atlantic coastal ecosystem not predicted by nutrient load. Journal of
Plankton Research 35: 1207-1219. http://dx.doi.org/10.1093/plankt/fbt070.
Trommer, G: Poxleitner, M: Lorenz, P; Bitzilekis, E: Gogaladze, A: Schultes, S: Stibor, H.
(2017).	Altered food-web dynamics under increased nitrogen load in phosphorus
deficient lakes. Aquat Sci 79: 1009-1021. http://dx.doi.org/10.1007/sQ0027-017-0551-2.
Tsui, MT: Finlav, JC: Nater, EA. (2008). Effects of stream water chemistry and tree species on
release and methylation of mercury during litter decomposition. Environ Sci Technol 42:
8692-8697. http://dx.doi.org/10.1021/es800956q.
June 2018
R-215
DRAFT: Do Not Cite or Quote

-------
Tu. L; Peng. Y; Chen. G; Hu. HL; Xiao. Y; Hu. TX; Liu. L; Tang. Y. (2015). Direct and indirect
effects of nitrogen additions on fine root decomposition in a subtropical bamboo forest.
Plant Soil 389: 273-288. http://dx.doi.org/10.1007/slllQ4-014-2353-9.
Tu. Q: Deng. Y; Yan. 0; Shen. L; Lin. L; He. Z; Wu. L; Van Nostrand. JD; Buzzard. V:
Michaletz. ST: Enquist. BJ; Weiser. MP: Kaspari. M; Waide. RB; Brown. JH; Zhou. J.
(2016). Biogeographic patterns of soil diazotrophic communities across six forests in the
North America. Mol Ecol 25: 2937-2948. http://dx.doi.Org/10.l 111/mec. 13651.
Tuccella. P; Curci. G: Grell. GA; Visconti. G: Crumevrolle. S: Schwarzenboeck. A: Mensah.
AA. (2015). A new chemistry option in WRF-Chem v. 3.4 for the simulation of direct
and indirect aerosol effects using VBS: evaluation against IMPACT-EUCAARI data.
Geoscientific Model Development 8: 2749-2776. http://dx.doi.org/10.5194/gmd-8-2749-
2015.
Tulloss. EM: Cadenasso. ML. (2016). The Effect of Nitrogen Deposition on Plant Performance
and Community Structure: Is It Life Stage Specific? PLoS ONE 11: e0156685.
http://dx.doi.org/10.1371/iournal.pone.0156685.
Tuomainen. U; Candolin. U. (2013). Environmental change and extended phenotypes: Does
eutrophication influence nest building in sticklebacks? Ethology 119: 503-510.
http://dx.doi.org/10. Ill 1/eth. 12088.
Turlapati. SA; Minocha. R; Bhiravarasa. PS: Tisa. LS: Thomas. WK; Minocha. SC. (2013).
Chronic N-amended soils exhibit an altered bacterial community structure in Harvard
Forest, MA, USA. FEMS Microbiol Ecol 83: 478-493. http://dx.doi.org/10.1111/1574-
6941.12009.
Turlapati. SA: Minocha. R; Long. S: Ramsdell. J: Minocha. SC. (2015). Oligotyping reveals
stronger relationship of organic soil bacterial community structure with N-amendments
and soil chemistry in comparison to that of mineral soil at Harvard Forest, MA, USA.
FMICB 6: 49. http://dx.doi.org/10.3389/fmicb.2015.00Q49.
Turner. EL: Paudel. B; Montagna. PA. (2015). Baseline nutrient dynamics in shallow well mixed
coastal lagoon with seasonal harmful algal blooms and hypoxia formation. Mar Pollut
Bull 96: 456-462. http://dx.doi.Org/10.1016/i.marpolbul.2015.05.005.
Turner. RE. (2011). Beneath the salt marsh canopy: Loss of soil strength with increasing nutrient
loads. Estuaries Coasts 34: 1084-1093. http://dx.doi.org/10.1007/sl2237-01Q-9341-v.
Turner. RE: Howes. BL; Teal. JM; Milan. CS: Swenson. EM: Goehringer-Toner. DP. (2009).
Salt marshes and eutrophi cation: An unsustainable outcome. Limnol Oceanogr 54: 1634-
1642. http://dx.doi.Org/10.4319/lo.2009.54.5.1634.
Turner. RE: Qureshi. N: Rabalais. NN: Dortch. O: Justic. D: Shaw. RF: Cope. J. (1998).
Fluctuating silicate:nitrate ratios and coastal plankton food webs. Proc Natl Acad Sci
USA 95: 13048-13051. http://dx.doi.org/10.1073/pnas.95.22.13048.
June 2018
R-216
DRAFT: Do Not Cite or Quote

-------
Turner. RE; Stanley. D; Brock. D; Pennock. J; Rabalais. NN. (2001). A comparison of
independent N-loading estimates for U.S. estuaries. In RA Valigura; RB Alexander; MS
Castro; TP Meyers; HW Paerl; PE Stacey; RE Turner (Eds.), Nitrogen Loading in
Coastal Water Bodies: An Atmospheric Perspective (Coastal and estuarine studies, 57
ed., pp. 107-118). Washington, DC.: American Geophysical Union.
http://dx.doi.org/10.1029/CE057pQ107.
Turner, RS; Cook, RB; Van Miegroet, H; Johnson, DW; Elwood, JW; Bricker, OP; Lindberg,
SE; Hornberger, GM. (1991). Watershed and lake processes affecting surface water acid-
base chemistry. In State-of-the-Science/Technology Report No 10. Washington, DC:
National Acid Precipitation Assessment Program.
Turney, CSM; Palmer, J; Maslin, MA; Hogg, A; Fogwill, CJ; Southon, J; Fenwick, P; Helle, G;
Wilmshurst, JM; Mcglone, M; Bronk Ramsey, C: Thomas, Z; Lipson, M; Beaven, B;
Jones, RT; Andrews, O: Hua, Q. (2018). Global peak in atmospheric radiocarbon
provides a potential definition for the onset of the anthropocene epoch in 1965. Sci Rep
8: 3293. http://dx.doi.org/10.1038/s41598-018-2097Q-5.
Turnipseed, AA; Huev, LG: Nemitz, E; Stickel, R; Higgs, J: Tanner, DJ; Slusher, PL: Sparks,
JP; Flocke, F; Guenther, A. (2006). Eddy covariance fluxes of peroxyacetyl nitrates
(PANs) and NOy to a coniferous forest. J Geophys Res Atmos 111: D09304.
http://dx.doi.org/10.1029/2005JD0Q6631.
Turnpenny, AWH; Sadler, K; Aston, RJ; Milner, AGP: Lynam, S. (1987). The fish populations
of some streams in Wales and northern England in relation to acidity and associated
factors. J Fish Biol 31: 415-434. http://dx.doi.Org/10.llll/i.1095-8649.1987.tb05246.x.
Turunen, J; Roulet, NT; Moore, TR; Richard, PJH. (2004). Nitrogen deposition and increased
carbon accumulation in ombrotrophic peatlands in eastern Canada. Global Biogeochem
Cycles 18. http://dx.doi.org/10.1029/2003GB0Q2154.
Twomev, LJ; Piehler, MF; Paerl, HW. (2005). Phytoplankton uptake of ammonium, nitrate and
urea in the Neuse River Estuary, NC, USA. Hydrobiologia 533: 123-134.
http://dx.doi. org/10.1007/s 10750-004-2403 -z.
Tyler, AC: Lambrinos, JG: Grosholz, ED. (2007). Nitrogen inputs promote the spread of an
invasive marsh grass. Ecol Appl 17: 1886-1898. http://dx.doi.org/10.1890/06-0822.1.
Tyler, RM; Brady, DC: Targett, TE. (2009). Temporal and spatial dynamics of diel-cycling
hypoxia in estuarine tributaries. Estuaries Coasts 32: 123-145.
http://dx.doi.org/10.1007/sl2237-008-91Q8-x.
U.S. EPA (U.S. Environmental Protection Agency). (1971). Air quality criteria for nitrogen
oxides [EPA Report], (AP-84). Washington DC: U.S. Environmental Protection Agency,
Air Pollution Control Office.
http://nepis.epa. gov/Exe/ZyPURL.cgi?Dockev=20013K3B.txt.
U.S. EPA (U.S. Environmental Protection Agency). (1973). Effects of sulfur oxides in the
atmosphere on vegetation; revised chapter 5 for "air quality criteria for sulfur oxides"
[EPA Report], (EPA-R3-73-030). Research Triangle Park, NC.
June 2018
R-217
DRAFT: Do Not Cite or Quote

-------
U.S. EPA (U.S. Environmental Protection Agency). (1982a). Air quality criteria for oxides of
nitrogen [EPA Report], (EPA/600/8-82-026). Research Triangle Park, NC: U.S.
Environmental Protection Agency, Environmental Criteria and Assessment Office.
U.S. EPA (U.S. Environmental Protection Agency). (1982b). Air quality criteria for particulate
matter and sulfur oxides (final, 1982) [EPA Report], (EPA 600/8-82/029a). Washington,
DC: Environmental Criteria and Assessment Office.
http://cfpub.epa. gov/ncea/cfm/recordisplav.cfm?deid=46205.
U.S. EPA (U.S. Environmental Protection Agency). (1982c). Review of the national ambient air
quality standards for particulate matter: Assessment of scientific and technical
information. OAQPS staff paper [EPA Report], (EPA/450/5-82/001). Research Triangle
Park, NC. http://nepis.epa.gov/exe/ZvPURL.cgi?Dockey=2000NH6N.txt.
U.S. EPA (U.S. Environmental Protection Agency). (1982d). Review of the national ambient air
quality standards for sulfur oxides: Assessment of scientific and technical information.
OAQPS staff paper - Final [EPA Report] (pp. 229). (EPA/450/5-82/007). Research
Triangle Park, NC: U.S. Environmental Protection Agency, Office of Air Quality
Planning and Standards, Strategies and Air Standards Division.
U.S. EPA (U.S. Environmental Protection Agency). (1984a). The acidic deposition phenomenon
and its effects: critical assessment review papers Volume I: atmospheric sciences [EPA
Report], (EPA/600/8-83/016AF). Washington, DC.
U.S. EPA (U.S. Environmental Protection Agency). (1984b). The acidic deposition phenomenon
and its effects: critical assessment review papers Volume II: effects sciences [EPA
Report], (EPA/600/8-83/016BF). Washington, DC.
U.S. EPA (U.S. Environmental Protection Agency). (1988). Ambient water quality criteria for
aluminum - 1988 [EPA Report], (EPA 440/5-86-008). Washington, DC.
https://nepis.epa.gov/Exe/ZvPDF.cgi/2000M5FC.PDF?Dockev=2000M5FC.PDF.
U.S. EPA (U.S. Environmental Protection Agency). (1993). Air quality criteria for oxides of
nitrogen (final, 1993) [EPA Report], (EPA/600/8-9l/049aF-cF). Washington, DC: U.S.
Environmental Protection Agency, Environmental Criteria and Assessment Office.
http://cfpub.epa. gov/ncea/cfm/recordisplav.cfm?deid=40179.
U.S. EPA (U.S. Environmental Protection Agency). (1995a). Acid deposition standard feasibility
study Report to Congress [EPA Report], (EPA 430-R-95-001a). Washington, DC.
U.S. EPA (U.S. Environmental Protection Agency). (1995b). Review of the national ambient air
quality standards for nitrogen dioxide: Assessment of scientific and technical information
[EPA Report], (EPA/452/R-95/005). Research Triangle Park, NC: U.S. Environmental
Protection Agency, Office of Air Quality, Planning and Standards.
http://nepis.epa. gov/exe/ZvPURL.cgi?Dockev=00002UBE.txt.
U.S. EPA (U.S. Environmental Protection Agency). (1996). Emissions Inventory Improvement
Program (EIIP) volume 6: Quality assurance procedures and DARS software [EPA
Report], (EPA-454/R-97-004). Washington, D.C.
June 2018
R-218
DRAFT: Do Not Cite or Quote

-------
U.S. EPA (U.S. Environmental Protection Agency). (1998a). Guidelines for ecological risk
assessment [EPA Report], (EPA/630/R-95/002F). Washington, DC: U.S. Environmental
Protection Agency, Risk Assessment Forum, https://www.epa.gov/risk/guidelines-
ecological-risk-assessment.
U.S. EPA (U.S. Environmental Protection Agency). (1998b). National strategy for the
development of regional nutrient criteria [EPA Report], (EPA/822/R-98/002).
Washington, DC: U.S. Environmental Protection Agency, Office of Water.
http://www.waterboards.ca.gov/sanfranciscobav/water issues/programs/planningtmdls/a
mendments/estuarineNNE/EP A%201998%20Nati onal%20 Strategy%20Document. pdf.
U.S. EPA (U.S. Environmental Protection Agency). (1999). The benefits and costs of the Clean
Air Act 1990 to 2010: EPA report to Congress [EPA Report], (EPA-410-R-99-001). U.S.
Environmental Protection Agency, Office of Air and Radiation, Office of Policy.
U.S. EPA (U.S. Environmental Protection Agency). (2001). Action plan for reducing, mitigating,
and controlling hypoxia in the northern Gulf of Mexico. Washington, DC: U.S.
Environmental Protection Agency, Office of Wetlands, Oceans, and Watersheds.
U.S. EPA (U.S. Environmental Protection Agency). (2003). Response of surface water chemistry
to the Clean Air Act Amendments of 1990 [EPA Report], (EPA 620/R-03/001). Research
Triangle Park, NC: U.S. Environmental Protection Agency, Office of Research and
Development, National Health and Environmental Effects Research Laboratory.
http://nepis.epa. gov/Exe/ZvPURL.cgi?Dockev=30003YXE.txt.
U.S. EPA (U.S. Environmental Protection Agency). (2004). Air quality criteria for particulate
matter [EPA Report], (EPA/600/P-99/002aF-bF). Research Triangle Park, NC: U.S.
Environmental Protection Agency, Office of Research and Development, National Center
for Environemntal Assessment- RTP Office.
http://cfpub.epa. gov/ncea/cfm/recordisplav.cfm?deid=87903.
U.S. EPA (U.S. Environmental Protection Agency). (2006a). Air quality criteria for ozone and
related photochemical oxidants [EPA Report], (EPA/600/R-05/004AF). Research
Triangle Park, NC: U.S. Environmental Protection Agency, Office of Research and
Development, National Center for Environmental Assessment-RTP Office.
http://cfpub.epa. gov/ncea/cfm/recordisplav.cfm?deid=149923.
U.S. EPA (U.S. Environmental Protection Agency). (2006b). Multi-agency critical loads
workshop: sulfur & nitrogen deposition effects on freshwater and terrestrial ecosystems
Final report [EPA Report], (EPA 68-W-03-02). Washington, DC.
U.S. EPA (U.S. Environmental Protection Agency). (2006c). Wadeable streams assessment: A
collaborative survey of the nation's streams [EPA Report], (EPA 841-B-06-002).
Washington, DC: U.S. Environmental Protection Agency, Office of Research and
Development, Office of Water, http://www.epa.gov/owow/streamsurvev.
U.S. EPA (U.S. Environmental Protection Agency). (2007). Integrated review plan for the
secondary national ambient air quality standards for nitrogen dioxide and sulfur dioxide
[EPA Report], (EPA-452/R-08-006). Washington, DC.
https://www3.epa.gov/ttn/naaqs/standards/no2so2sec/data/20071220 final review plan.p
df.
June 2018	R-219	DRAFT: Do Not Cite or Quote

-------
U.S. EPA (U.S. Environmental Protection Agency). (2008a). Integrated science assessment for
oxides of nitrogen and sulfur: Ecological criteria [EPA Report], (EPA/600/R-08/082F).
Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Research
and Development, National Center for Environmental Assessment- RTP Division.
http://cfpub.epa. gov/ncea/cfm/recordisplav.cfm?deid=201485.
U.S. EPA (U.S. Environmental Protection Agency). (2008b). Integrated science assessment for
oxides of nitrogen Health criteria (Final report, 2008) [EPA Report], (EPA/600/R-
08/071). Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of
Research and Development, National Center for Environmental Assessment- RTP
Division, http://cfpub.epa.gov/ncea/cfm/recordisplay.cfm?deid= 194645.
U.S. EPA (U.S. Environmental Protection Agency). (2008c). Integrated science assessment for
sulfur oxides: Health criteria [EPA Report], (EPA/600/R-08/047F). Research Triangle
Park, NC: U.S. Environmental Protection Agency, Office of Research and Development,
National Center for Environmental Assessment- RTP.
http://cfpub.epa. gov/ncea/cfm/recordisplav.cfm?deid=198843.
U.S. EPA (U.S. Environmental Protection Agency). (2009a). Integrated science assessment for
particulate matter [EPA Report], (EPA/600/R-08/139F). Research Triangle Park, NC:
U.S. Environmental Protection Agency, Office of Research and Development, National
Center for Environmental Assessment- RTP Division.
http://cfpub.epa. gov/ncea/cfm/recordisplav.cfm?deid=216546.
U.S. EPA (U.S. Environmental Protection Agency). (2009b). National lakes assessment: A
collaborative survey of the nations lakes. (EPA/84 l/R-09/001). Washington, DC: U.S.
Environmental Protection Agency, Office of Research and Development, Office of
Water, https://www.epa.gov/sites/production/files/2013-
11/documents/nla newlowres fullrpt.pdf.
U.S. EPA (U.S. Environmental Protection Agency). (2009c). Risk and exposure assessment for
review of the secondary National Ambient Air Quality Standards for oxides of nitrogen
and oxides of sulfur [EPA Report], (EPA/452/R-09/008A). Research Triangle Park, NC:
U.S. Environmental Protection Agency, Office of Air Quality Planning and Standards,
Health and Environmental Impacts Division.
U.S. EPA (U.S. Environmental Protection Agency). (2010). Establishing the allocations for the
basin-jurisdictions [EPA Report], In Chesapeake Bay total maximum daily load for
nitrogen, phosphorus and sediment (pp. 1-53). Washington, DC.
https://www.epa.gov/chesapeake-bav-tmdl/chesapeake-bav-tmdl-document.
U.S. EPA (U.S. Environmental Protection Agency). (201 la). Policy assessment for the review of
the secondary national ambient air quality standards for oxides of nitrogen and oxides of
sulfur [EPA Report], (EPA-452/R-1 l-004a). Washington, DC.
http://www3.epa.gov/ttn/naaqs/standards/no2so2sec/data/20110114pamain.pdf.
June 2018
R-220
DRAFT: Do Not Cite or Quote

-------
U.S. EPA (U.S. Environmental Protection Agency). (2011b). Revised technical support
document: National-scale assessment of mercury risk to populations with high
consumption of self-caught freshwater fish. In support of the appropriate and necessary
finding for coal- and oil-fired electric generating units. (EPA-452/R-11-009;
NTIS/10450099). Research Triangle Park, NC: U.S. Environmental Protection Agency,
Office of Air Quality Planning and Standards, Health and Environmental Impacts
Division.
U.S. EPA (U.S. Environmental Protection Agency). (2012a). Air trends: Sulfur dioxide.
Available online at https://www.epa.gov/air-trends/sulfur-dioxide-trends (accessed
U.S. EPA (U.S. Environmental Protection Agency). (2012b). National coastal condition report
IV [EPA Report], (EPA-842-R-10-003). Washington, DC: U.S. Environmental Protection
Agency, Office of Research and Development/Office of Water.
http://water.epa.gOv/tvpe/oceb/assessmonitor/nccr/upload/0 NCCR 4 Report 508 book
marks.pdf.
U.S. EPA (U.S. Environmental Protection Agency). (2012c). Technical support document for
U.S. EPA's proposed rule for numeric nutrient criteria for Floridas estuaries, coastal
waters, and south Florida inland flowing waters, Volume 1: Estuaries (pp. 1-365).
Washington, DC: U.S. Environmental Protection Agency, Office of Water.
U.S. EPA (U.S. Environmental Protection Agency). (2013a). Aquatic life ambient water quality
criteria for ammonia Freshwater [EPA Report], (EPA 822-R-13-001). Washington, DC:
U.S. Environmental Protection Agency, Office of Water, Office of Science and
Technology, https://www.epa.gov/wqc/aquatic-life-criteria-ammonia.
U.S. EPA (U.S. Environmental Protection Agency). (2013b). Integrated science assessment for
lead [EPA Report], (EPA/600/R-10/075F). Research Triangle Park, NC: U.S.
Environmental Protection Agency, National Center for Environmental Assessment.
http://cfpub.epa.gov/ncea/cfm/recordisplav.cfm?deid=255721.
U.S. EPA (U.S. Environmental Protection Agency). (2013c). Integrated science assessment for
ozone and related photochemical oxidants [EPA Report], (EPA/600/R-10/076F).
Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Research
and Development, National Center for Environmental Assessment-RTP Division.
http://cfpub.epa. gov/ncea/isa/recordisplav.cfm?deid=247492.
U.S. EPA (U.S. Environmental Protection Agency). (2015a). Drinking water health advisory for
the cyanobacterial microcystin toxins [EPA Report], (EPA- 820R15100). Washington,
DC: U.S. Environmental Protection Agency, Office of Water.
http://www2.epa.gov/sites/production/files/2015-06/documents/microcvstins-report-
2015.pdf.
U.S. EPA (U.S. Environmental Protection Agency). (2015b). Drinking water health advisory for
the cyanobacterial toxin cylindrospermopsin [EPA Report], (EPA- 820R15101).
Washington, DC: U.S. Environmental Protection Agency, Office of Water.
http://www2.epa.gov/sites/production/files/2015-06/documents/cvlindrospermopsin-
report-2015.pdf.
June 2018	R-221	DRAFT: Do Not Cite or Quote

-------
U.S. EPA (U.S. Environmental Protection Agency). (2015c). Mississippi River/Gulf of Mexico
watershed nutrient task force 2015 report to congress [EPA Report], Washington, DC:
U.S. Environmental Protection Agency, National Center for Environmental Asseessment.
http://www2.epa.gov/sites/production/files/2015-
10/documents/htf report to congress final - 10.1.15.pdf.
U.S. EPA (U.S. Environmental Protection Agency). (2015d). National ecosystem services
classification system (NESCS): Framework design and policy application (final report)
[EPA Report], (EPA-800-R-15-002). Washington, DC: U.S. Environmental Protection
Agency, Office of Water, Office of Research and Development.
https://www.epa.gov/eco-research/national-ecosvstem-services-classification-svstem-
framework-design-and-policv.
U.S. EPA (U.S. Environmental Protection Agency). (2015e). Preamble to the integrated science
assessments [EPA Report], (EPA/600/R-15/067). Research Triangle Park, NC: U.S.
Environmental Protection Agency, Office of Research and Development, National Center
for Environmental Assessment, RTP Division.
https://cfpub.epa.gOv/ncea/i sa/recordisplav.cfm?deid=310244.
U.S. EPA (U.S. Environmental Protection Agency). (2016a). 2014 National Emissions Inventory
(NEI) data. Washington, DC. Retrieved from https://www.epa.gov/air-emissions-
inventories/2014-national-emissions-inventory-nei-data
U.S. EPA (U.S. Environmental Protection Agency). (2016b). 2014 National Emissions
Inventory, version 1 technical support document (Draft). Research Triangle Park, NC.
U.S. EPA (U.S. Environmental Protection Agency). (2016c). Clean Air Status and Trends
Network (CASTNET) quality assurance aeport Fourth Quarter 2016 with 2016 Annual
Summary [EPA Report], Washington, D.C.
U.S. EPA (U.S. Environmental Protection Agency). (2016d). Integrated review plan for the
national ambient air quality standards for particulate matter [EPA Report], (EPA-452/R-
16-005). Research Triangle Park, NC.
https://vosemite.epa.gOv/sab/sabproduct.nsf/0/eb862b233fbd0cde85257dda004fcb8clQpe
nDocument&TableRow=2.0.
U.S. EPA (U.S. Environmental Protection Agency). (2016e). Integrated science assessment
(ISA) for sulfur oxides health criteria (Second external review draft) [EPA Report],
(EPA/600/R-16/351). Washington, DC.
https://cfpub.epa.gOv/ncea/i sa/recordisplav.cfm?deid=326450.
U.S. EPA (U.S. Environmental Protection Agency). (2016f). Integrated science assessment for
oxides of nitrogen-Health Criteria (final report) [EPA Report], (EPA/600/R-15/068).
Research Triangle Park, NC: U.S. Environmental Protection Agency, Office of Research
and Development, National Center for Environmental Assessment.
http://ofmpub.epa.gov/eims/eimscomm.getfile7p download id=526855.
June 2018
R-222
DRAFT: Do Not Cite or Quote

-------
U.S. EPA (U.S. Environmental Protection Agency). (2016g). National coastal condition
assessment 2010. (EPA 841-R-15-006). Washington, DC: U.S. Environmental Protection
Agency, Office of Water, Office of Research and Development.
https://www.epa.gov/national-aquatic-resource-survevs/ncca.
U.S. EPA (U.S. Environmental Protection Agency). (2016h). National lakes assessment 2012: A
collaborative survey of lakes in the United States [EPA Report], (EPA 841-R-16-113).
Washington, D.C. https://nationallakesassessment.epa.gov/.
U.S. EPA (U.S. Environmental Protection Agency). (2016i). National rivers and streams
assessment 2008-2009: A collaborative survey. (EPA/841/R-16/007). Washington, DC:
U.S. Environmental Protection Agency, Office of Water, Office of Research and
Development, https://www.epa.gov/national-aquatic-resource-surveys/national-rivers-
and-streams-assessment-2008-2009-report.
U.S. EPA (U.S. Environmental Protection Agency). (2016j). National wetland condition
assessment 2011: A collaborative survey of the nation's wetlands (pp. 1-105). (EPA-843-
R-l5-005). Washington, DC. https://www.epa.gov/national-aquatic-resource-
survevs/national-wetland-condition-assessment-2011 -report.
U.S. EPA (U.S. Environmental Protection Agency). (2016k). US EPAs report on the
environment (ROE). Available online at https://cfpub.epa.gov/roe/index.cfm (accessed
U.S. EPA (U.S. Environmental Protection Agency). (2017a). EPA's report on the environment
(ROE). Available online at https://cfpub.epa. gov/roe/index.cfm (accessed January 29,
2018).
U.S. EPA (U.S. Environmental Protection Agency). (2017b). Integrated review plan for the
secondary national ambient air quality standards for ecological effects of oxides of
nitrogen, oxides of sulfur and particulate matter [EPA Report], (EPA-452/R-17-002).
Research Triangle Park, NC.
U.S. EPA (U.S. Environmental Protection Agency). (2017c). Integrated science assessment for
sulfur oxides: Health criteria [EPA Report], (EPA/600/R-17/451). Research Triangle
Park, NC: U.S. Environmental Protection Agency, Office of Research and Development,
National Center for Environmental Assessment- RTP.
https://www.epa.gov/isa/integrated-science-assessment-isa-sulfur-oxides-health-criteria.
U.S. EPA (U.S. Environmental Protection Agency). (2017d). Mississippi River/Gulf of Mexico
Watershed Nutrient Task Force 2017 Report to Congress. Available online at (accessed
U.S. EPA (U.S. Environmental Protection Agency). (2017e). Northern Gulf of Mexico hypoxic
zone [Website], Washington, DC. Retrieved from https://www.epa.gov/ms-htf/northern-
gulf-mexico-hypoxic-zone
U.S. Global Change Research Program. (2009). Global climate change impacts in the United
States: A state of knowledge report. In Global climate change impacts in the United
States : a state of knowledge report. New York, NY: Cambridge University Press.
https://downloads.globalchange.gov/usimpacts/pdfs/climate-impacts-report.pdf.
Uchytil RJ. (1991). Abies balsamea. In: Fire Effects Information System. Available online at
http://www.fs.fed.us/database/feis/plants/tree/abibal/all.html (accessed March 17, 2016).
June 2018
R-223
DRAFT: Do Not Cite or Quote

-------
Ukonmaanaho. L; Starr. M; Kantola. M; Lauren. A. ri; Piispanen. J; Pietila. H; Peramaki. P;
Merila. P; Fritze. H; Tuomivirta. T; Heikkinen. J; Makinen. J; Nieminen. TM. (2016).
Impacts of forest harvesting on mobilization of Hg and MeHg in drained peatland forests
on black schist or felsic bedrock. Environ Monit Assess 188: 228.
http://dx.doi.org/10.1007/slQ661-016-5210-x.
UNEP (United Nations Environment Programme). (2007). Global environment outlook 4 :
Environment for development. In Global Environment Outlook 4 : Environment for
Development. Nairobi, Kenya.
University of Kentucky. (2015). Fraxinus americana (white ash) olive family (Oleaceae) [Fact
Sheet], Lexington, KY: Univeristy of Kentucky, College of Agriculture.
http://www.ukv.edu/hort/sites/www.ukv.edu.hort/files/pages-
attachments/FRAXIAMprint. pdf.
University of Michigan. (2015). Native American ethnobotany. Available online at
http://herb.umd.umich.edu/ (accessed March 17, 2016).
Urakawa. H; Martens-Habbena. W: Huguet. C: de La Torre. JR; Ingalls. AE; Devol. AH: Stahl.
DA. (2014). Ammonia availability shapes the seasonal distribution and activity of
archaeal and bacterial ammonia oxidizers in the Puget Sound Estuary. Limnol Oceanogr
59: 1321-1335. http://dx.doi.Org/10.4319/lo.2014.59.4.1321.
Urban. NR; Brezonik. PL: Baker. LA: Sherman. LA. (1994). Sulfate reduction and diffusion in
sediments of Little Rock Lake, Wisconsin. Limnol Oceanogr 39: 797-815.
http://dx.doi.Org/10.4319/lo.1994.39.4.0797.
USD A (U.S. Department of Agriculture). (2010). Census of horticultural specialties 2009
Volume 3. Special studies. Part 3. (AC-07-SS-3). Washington, DC.
http://www.agcensus.usda.gov/Publications/2007/Qnline Highlights/Census of Horticul
ture Specialties/HORTIC.pdf.
USD A (U.S. Department of Agriculture). (2015a). Natural resources conservation service plant
database: American Eelgrass. Available online at
http://plants.usda.gov/core/profile?svmbol=vaam3 (accessed
USD A (U.S. Department of Agriculture). (2015b). Natural Resources Conservation Service
PLANTS database. Available online at http://plants.usda.gov/iava/ (accessed
USD A (U.S. Department of Agriculture, Forest Service). (1990a). Fraxinus americana L. white
ash. In RM Burns; BH Honkala (Eds.), Silvics of North America: 1 Conifers; 2
Hardwoods Agriculture Handbook 654. Washington, DC: U.S. Department of
Agriculture, U.S. Forest Service.
http://www.na.fs.fed.us/spfo/pubs/silvics manual/table of contents.htm.
USD A (U.S. Department of Agriculture, Forest Service). (1990b). Volume 1: Conifers: Abies
balsamea (L.) mill. Balsam fir. In RM Burns; BH Honkala (Eds.), Silvics of North
America (pp. 26-47). Washington, DC: U.S. Department of Agriculture, U.S. Forest
Service, http://www.na.fs.fed.us/spfo/pubs/silvics manual/table of contents.htm.
June 2018
R-224
DRAFT: Do Not Cite or Quote

-------
USDANRCS (U.S. Department of Agriculture Natural Resources Conservation Service). White
ash. Plant guide fact sheet [Fact Sheet], Washington, DC: U.S. Department of
Agriculture, Natural Resources Conservation Service.
http://plants.usda.gov/plantguide/pdf/cs fram2.pdf.
USFWS (U.S. Fish and Wildlife Service). (1988). The ecology of Tampa Bay, Florida: An
estuarine profile. Washington, DC: U.S. Department of Interior, U.S. Fish and Wildlife
Service, http://www.nwrc.usgs.gov/techrpt/85-7-18.pdf.
USFWS (U.S. Fish and Wildlife Service). (1990). An ecological characterization of the Tampa
Bay watershed. (Biological Report 90(20)). Washington, DC: U.S. Department of the
Interior, U.S. Fish and Wildlife Service.
USFWS (U.S. Fish and Wildlife Service). (2015). Green sea turtle (Chelonia mydas) [Fact
Sheet], Arlington, VA.
http://www.fws.gov/northflorida/seaturtles/turtle%20factsheets/green-sea-turtle.htm.
USGS (U.S. Geological Survey). (2011). Integrating science and resource management in Tampa
Bay, Florida. Reston, VA: U.S. Department of Interior, U.S. Geological Survey.
http://pubs.usgs.gov/circ/1348/.
USGS US Department of the Interior (United States Geological Survey US Department of the
Interior). (2015). Estuarine water quality in parks of the Northeast Coastal and Barrier
Network: Vital signs estuarine nutrient-enrichment monitoring, 2006-11. (Natural
Resource Technical Report NPS/NCBN/NRR - 2015/902). Ft Collins, CO: National Park
Service.
Uzu. G: Sobanska. S: Sarret G: Munoz. M; Dumat C. (2010). Foliar lead uptake by lettuce
exposed to atmospheric fallouts. Environ Sci Technol 44: 1036-1042.
http://dx.doi.org/10.1021/es9Q2190u.
Vadeboncoeur, MA; Hamburg, SP; Pryor, D. (2010). Modeled nitrogen loading to Narragansett
Bay: 1850 to 2015. Estuaries Coasts 33: 1113-1127. http://dx.doi.org/10.10Q7/sl2237-
010-9320-3.
Vadeboncoeur. Y; Peterson. G: Vander Zanden. MJ; Kalff J. (2008). Benthic algal production
across lake size gradients: Interactions among morphometry, nutrients, and light. Ecology
89: 2542-2552. http://dx.doi.Org/10.1890/07-1058.l.
Valiela. I: Foreman. K; LaMontagne. M; Hersh. D; Costa. J: Peckol. P; DeMeo-Andreson. B;
D'Avanzo. C: Babione. M; Sham. C. -H; Brawlev. J: Laitha. K. (1992). Couplings of
watersheds and coastal waters: Sources and consequences of nutrient enrichment in
WaquoitBay, Massachusetts. Estuaries Coasts 15: 443-457.
http://dx.doi.org/10.2307/1352389.
Valiela. I: Owens. C: Elmstrom. E; Lloret. J. (2016). Eutrophication of Cape Cod estuaries:
Effect of decadal changes in global-driven atmospheric and local-scale wastewater
nutrient loads. Mar Pollut Bull 110: 309-315.
http://dx.doi.Org/10.1016/i.marpolbul.2016.06.047.
Valinia. S: Futter. MN; Cosby. BJ; Rosen. P; Folster. J. (2015). Simple models to estimate
historical and recent changes of total organic carbon concentrations in lakes. Environ Sci
Technol 49: 386-394. http://dx.doi.org/10.1021/es503170r.
June 2018
R-225
DRAFT: Do Not Cite or Quote

-------
Vallack, HW; Leronni, V; Metcalfe. DB; Hogberg, P; Ineson, P; Subke, J. -A. (2012).
Application of nitrogen fertilizer to a boreal pine forest has a negative impact on the
respiration of ectomycorrhizal hyphae. Plant Soil 352: 405-417.
http://dx.doi.org/10.1007/sl 1104-011-1005-6.
Vallano. DM: Selmants. PC: Zavaleta. ES. (2012). Simulated nitrogen deposition enhances the
performance of an exotic grass relative to native serpentine grassland competitors. Plant
Ecol 213: 1015-1026. http://dx.doi.org/10.1007/sll258-012-0Q61-l.
Vallano. DM: Sparks. JP. (2008). Quantifying foliar uptake of gaseous nitrogen dioxide using
enriched foliar 815N values. NewPhytol 177: 946-955. http://dx.doi.Org/10.l 11 l/j.1469-
8137.2007.02311.x.
Valliere. JM; Allen. EB. (2016a). Interactive effects of nitrogen deposition and drought-stress on
plant-soil feedbacks of Artemisia californica seedlings. Plant Soil 403: 277-290.
http://dx.doi.org/10.1007/slll04-015-2776-v.
Valliere. JM: Allen. EB. (2016b). Nitrogen enrichment contributes to positive responses to soil
microbial communities in three invasive plant species. Biological Invasions 18: 2349-
2364. http://dx.doi.org/10.1007/slQ530-016-1166-6.
Valois, A; Keller, WB; Ramcharan, C. (2010). Abiotic and biotic processes in lakes recovering
from acidification: The relative roles of metal toxicity and fish predation as barriers to
zooplankton re-establishment. FreshwBiol 55: 2585-2597.
http://dx.doi.org/10.111 l/i.l365-2427.2010.02488.x.
Valois, AE; Keller, WB; Ramcharan, CW. (2011). Recovery in a multiple stressor environment:
Using the reference condition approach to examine zooplankton community change along
opposing gradients. Journal of Plankton Research 33: 1417-1429.
http://dx.doi.org/10.1093/plankt/fbr036.
Vamstad, MS: Rotenberry, JT. (2010). Effects of fire on vegetation and small mammal
communities in a Mojave Desert Joshua tree woodland. J Arid Environ 74: 1309-1318.
http://dx.doi.Org/10.1016/i.iaridenv.2010.04.002.
Van Breemen, N: Bover, EW: Goodale, CL; Jaworski, NA; Paustian, K; Seitzinger, SP; Laitha,
K; Mayer, B; Van Dam, D; Howarth, RW; Nadelhoffer, KJ; Eve, M; Billen, G. (2002).
Where did all the nitrogen go? Fate of nitrogen inputs to large watersheds in the
northeastern USA. Biogeochemistry 57/58: 267-293.
Van Breemen, N: Driscoll, CT; Mulder, J. (1984). Acidic deposition and internal proton sources
in acidification of soils and waters. Nature LOND: 599-604.
Van Damme, M; Clarisse, L; Heald, CL: Hurtmans, D; Ngadi, Y; Clerbaux, C: Dolman, AJ;
Erisman, JW: Coheur, PF. (2014). Global distributions, time series and error
characterization of atmospheric ammonia (NH3) from IASI satellite observations. Atmos
Chem Phys 14: 2905-2922. http://dx.doi.org/10.5194/acp-14-2905-2014.
van Den Berg, LJL; Jones, L; Sheppard, LJ; Smart, SM; Bobbink, R; Pise, NB; Ashmore, MR.
(2016). Evidence for differential effects of reduced and oxidised nitrogen deposition on
vegetation independent of nitrogen load. Environ Pollut 208: 890-897.
http://dx.doi.Org/10.1016/i.envpol.2015.09.017.
June 2018
R-226
DRAFT: Do Not Cite or Quote

-------
van Den Elzen, E, va; Kox, MAR; Harpenslager, SF; Hensgens, G; Fritz. C; Jetten, MSM;
Ettwig, KF; Lamers. LPM. (2017). Symbiosis revisited: phosphorus and acid buffering
stimulate N-2 fixation but not Sphagnum growth. Biogeosciences 14: 1111-1122.
http://dx.doi.org/10.5194/bg-14-l 111-2017.
Van Per Heiiden, MGA; Klironomos. JN; Ursic. M: Moutoglis. P; Streitwolf-Engel. R: Boiler.
T: Wiemken. A: Sanders. IR. (1998). Mycorrhizal fungal diversity determines plant
biodiversity, ecosystem variability and productivity. Nature 396: 69-72.
Van Per Heiiden. MGA: Verkade. S: De Bruin. SJ. (2008). Mycorrhizal fungi reduce the
negative effects of nitrogen enrichment on plant community structure in dune grassland.
Global Change Biol 14: 2626-2635. http://dx.doi.Org/10.llll/i.1365-2486.2008.01691.x.
van Diepen. LT; Frev. SD; Landis. EA; Morrison. EW: Pringle. A. (2017). Fungi exposed to
chronic nitrogen enrichment are less able to decay leaf litter. Ecology 98: 5-11.
http://dx.doi.org/10.1002/ecv.1635.
van Diepen. LTA; Entwistle. EM: Zak. DR. (2013). Chronic nitrogen deposition and the
composition of active arbuscular mycorrhizal fungi. Appl Soil Ecol 72: 62-68.
http://dx.doi.Org/10.1016/i.apsoil.2013.05.012.
van Diepen, LTA; Frev, SD: Sthultz, CM; Morrison, EW: Minocha, R; Pringle, A. (2015).
Changes in litter quality caused by simulated nitrogen deposition reinforce the N-induced
suppression of litter decay. Ecosphere 6: 1-16. http://dx.doi.Org/10.1890/ES15-00262.l.
van Diepen, LTA: Lilleskov, EA: Pregitzer, KS. (2011). Simulated nitrogen deposition affects
community structure of arbuscular mycorrhizal fungi in northern hardwood forests. Mol
Ecol 20: 799-811. http://dx.doi.org/10.111 l/i.l365-294X.2010.04969.x.
van Diepen, LTA: Lilleskov, EA: Pregitzer, KS: Miller, RM. (2010). Simulated nitrogen
deposition causes a decline of intra- and extraradical abundance of arbuscular
mycorrhizal fungi and changes in microbial community structure in northern hardwood
forests. Ecosystems 13: 683-695. http://dx.doi.org/10.1007/sl0021-010-9347-Q.
van Dobben, H; de Vries, W. (2010). Relation between forest vegetation, atmospheric deposition
and site conditions at regional and European scales. Environ Pollut 158: 921-933.
http://dx.doi.Org/10.1016/i.envpol.2009.09.015.
van Duinen, GA; Brock, AMT; Kuper, JT; Peeters, TMJ; Esselink, J. (2004). Do raised bog
restoration measures rehabilitate aquatic fauna diversity? A comparative study between
pristine, degraded, and rewetted raised bogs. In J Paivanen (Ed.), Wise Use of Peatlands:
Proceedings of the 12th International Peat Congress (pp. 399-405). Jyvaskyla, Finland:
International Peat Society.
van Egmond, K; Bresser, T; Bouwman, L. (2002). The European nitrogen case. Ambio 31: 72-
78. http://dx.doi.Org/10.1579/0044-7447-31.2.72.
Van Grinsven, HJM; Holland, M; Jacobsen, BH; Klimont, Z; Sutton, MA; Willems, WJ. (2013).
Costs and benefits of nitrogen for Europe and implications for mitigation. Environ Sci
Technol 47: 3571-3579. http://dx.doi.org/10.1021/es303804g.
June 2018
R-227
DRAFT: Do Not Cite or Quote

-------
van Herk. CM. (2001). Bark pH and susceptibility to toxic air pollutants as independent causes
of changes in epiphytic lichen composition in space and time. Lichenologist 33: 419-441.
http://dx.doi.ore/10.1006/lich.2001.0337.
Van Miegroet H; Creed. IF: Nicholas. NS: Tarboton. DG: Webster. KL; Shubzda. J: Robinson.
B; Smoot. JL; Johnson. DW: Lindberg. SE; Lovett. G: Nodvin. S: Moore. S. (2001). Is
there synchronicity in nitrogen input and output fluxes at the Noland Divide watershed, a
small N-saturated forested catchment in the Great Smoky Mountains National Park?
ScientificWorldJournal 1: 480-492. http://dx.doi.Org/10.l 100/tsw.2001.384.
Van Miegroet. H; Moore. PT; Tewksbury. CE; Nicholas. NS. (2007). Carbon sources and sinks
in high-elevation spruce-fir forests of the southeastern US. For Ecol Manage 238: 249-
260. http://dx.doi.Org/10.1016/i.foreco.2006.10.020.
van Wiinen. HJ; Bakker. JP. (1999). Nitrogen and phosphorus limitation in a coastal barrier salt
marsh: The implications for vegetation succession. J Ecol 87: 265-272.
http://dx.doi.Org/10.1046/i.1365-2745.1999.00349.x.
Vanguelova. EI; Hirano. Y; Eldhuset. TP: Sas-Paszt. L; Bakker. MR: Puttsepp. U: Brunner. I;
Lohmus. K; Godbold. D. (2007). Tree fine root Ca/Al molar ratio - Indicator of A1 and
acidity stress. Plant Biosyst 141: 460-480.
http://dx.doi.org/10.1080/112635007Q1626192.
Vankoughnett. MR; Henry. HAL. (2014). Soil freezing and N deposition: Transient vs multi-
year effects on plant productivity and relative species abundance. New Phytol 202: 1277-
1285. http://dx.doi.org/10.Ill 1/nph. 12734.
Vaguer-Sunyer. R; Duarte. CM. (2011). Temperature effects on oxygen thresholds for hypoxia in
marine benthic organisms. Global Change Biol 17: 1788-1797.
http://dx.doi.org/10.1111/i. 1365-2486.2010.02343 ,x.
Veblen, TT; Hadley, KS; Nel, EM; Kitzberger, T; Reid, M; Villalba, R. (1994). Disturbance
regime and disturbance interactions in a Rocky Mountain subalpine forest. J Ecol 82:
125-135. http://dx.doi.org/10.2307/2261392.
Vega Thurber. RL; Burkepile. DE; Fuchs. C; Shantz. AA; McMinds. R; Zaneveld. JR. (2014).
Chronic nutrient enrichment increases prevalence and severity of coral disease and
bleaching. Global Change Biol 20: 544-554. http://dx.doi.org/10. Ill 1/gcb. 12450.
Verburg. PSJ; Young. AC; Stevenson. BA; Glanzmann. I; Arnone. JA. Ill; Marion. GM;
Holmes. C; Nowak. RS. (2013). Do increased summer precipitation and N deposition
alter fine root dynamics in a Mojave Desert ecosystem? Global Change Biol 19: 948-956.
http://dx.doi.org/10. Ill 1/gcb. 12082.
Verheven. K; Baeten. L; De Frenne. P; Bernhardt-Roemermann. M; Brunet. J; Cornelis. J;
Decocq. G; Dierschke. H; Eriksson. O; Hedl. R; Heinken. T; Hermy. M; Hommel. P;
Kirby, K; Naaf, T; Peterken, G; Petrik, P; Pfadenhauer, J; Van Calster, H; Walther, G, -
R; Wulf. M; Verstraeten. G. (2012). Driving factors behind the eutrophication signal in
understorey plant communities of deciduous temperate forests. J Ecol 100: 352-365.
http://dx.doi.org/10.1111/i. 1365-2745.2011.01928.x.
June 2018
R-228
DRAFT: Do Not Cite or Quote

-------
Verhoeven. JTA; Bel tin an. B; Dorland. E; Robat SA; Bobbink. R. (2011). Differential effects of
ammonium and nitrate deposition on fen phanerogams and bryophytes. Appl Veg Sci 14:
149-157. http://dx.doi.org/10.111 l/i.l654-109X.2010.01113.x.
Verity. PG: Borkman. DG. (2010). A decade of change in the Skidaway River estuary. III.
Plankton. Estuaries Coasts 33: 513-540. http://dx.doi.org/10.1007/sl2237-009-92Q8-2.
Venn a. V: Rico-Martinez. R; Kotra. N: Retinoids. C: Liu. J: Snell. TW: Weber. RJ. (2013).
Estimating the toxicity of ambient fine aerosols using freshwater rotifer Brachionus
calyciflorus (Rotifera: Monogononta). Environ Pollut 182: 379-384.
http://dx.doi.Org/10.1016/i.envpol.2013.07.037.
Vermaat. J. anE; Harmsen. J: Hellmann. FA: van Per Geest. HG: de Klein. JJM; Kosten. S:
Smolders. AJP; Verhoeven. J. osTA; Mes. R. onG; Ouboter. M. (2016). Annual sulfate
budgets for Dutch lowland peat polders: The soil is a major sulfate source through peat
and pyrite oxidation. J Hydrol 533: 515-522.
http://dx.doi.Org/10.1016/i.ihydrol.2015.12.038.
Vernile. P; Tutino. M; Bari. G: Amodio. M; Spagnuolo. M; de Gennaro. G: de Lillo. E. (2013).
Particulate matter toxicity evaluation using bioindicators and comet assay. Aerosol Air
Qual Res 13: 172-178. http://dx.doi.org/10.4209/aaar.2012.07.0173.
Vertucci. FA: Corn. PS. (1996). Evaluation of episodic acidification and amphibian declines in
the Rocky Mountains. Ecol Appl 6: 449-457.
Vertucci. FA: Eilers. JM. (1993). Issues in monitoring wilderness lake chemistry: A case study in
the Sawtooth Mountains, Idaho. Environ Monit Assess 28: 277-294.
http://dx.doi.org/10.1007/BF0Q545771.
Vet. R; Artz. RS: Carou. S: Shaw. M; Ro. C. -U: Aas. W: Baker. A: Bowersox. VC: Dentener. F;
Galv-Lacaux. C: Hou. A: Pienaar. JJ; Gillett. R; Cristina Forti. M; Gromov. S: Hara. H;
Khodzher, T; Mahowald, NM; Nickovic, S; Rao, PSP; Reid, NW. (2014). A global
assessment of precipitation chemistry and deposition of sulfur, nitrogen, sea salt, base
cations, organic acids, acidity and pH, and phosphorus. Atmos Environ 93: 3-100.
http://dx.doi.Org/10.1016/i.atmosenv.2013.10.060.
Viard, B; Pihan, F; Promeyrat, S; Pihan, JC. (2004). Integrated assessment of heavy metal (Pb,
Zn, Cd) highway pollution: bioaccumulation in soil, graminaceae and land snails.
Chemosphere 55: 1349-1314. http://dx.doi.Org/10.1016/i.chemosphere.2004.01.003.
Vicca. S: Luvssaert. S: Penuelas. J: Campioli. M; Chapin. FS. Ill: Ciais. P; Heinemever. A:
Hogberg, P; Kutsch, WL; Law, BE; Malhi, Y; Papale, D; Piao, SL; Reichstein, M;
Schulze. ED: Janssens. IA. (2012). Fertile forests produce biomass more efficiently. Ecol
Lett 15: 520-526. http://dx.doi.org/10.111 l/i.l461-0248.2012.01775.x.
Vilmi, A: Karialainen, SM; Hellsten, S: Heino, J. (2016). Bioassessment in a metacommunity
context: Are diatom communities structured solely by species sorting? Ecol Indicat 62:
86-94. http://dx.doi.org/10.1016/i.ecolind.2015.11.043.
VIMS (Virginia Institute of Marine Science). (2016). SAV area in Chesapeake Bay and coastal
bays. Available online at http://web.vims.edu/bio/sav/index.html (accessed
June 2018
R-229
DRAFT: Do Not Cite or Quote

-------
Vinebrooke. RD; Maclennan. MM; Bartrons. M; Zettel. JP. (2014). Missing effects of
anthropogenic nutrient deposition on sentinel alpine ecosystems. Global Change Biol 20:
2173-2182. http://dx.doi.org/10.1111/gcb. 12484.
Vinebrooke. RD: Turner. MA: Findlav. PL: Paterson. MJ. (2009). A stressor-independent test
for biodiversity-ecosystem function relationships during a 23-year whole-lake
experiment. Can J Fish Aquat Sci 66: 1903-1909. http://dx.doi.Org/10.l 139/F09-120.
Vitousek. PM; Aber. JD; Howarth. RW: Likens. GE; Matson. PA: Schindler. DW: Schlesinger.
WH: Tilman. DG. (1997). Human alteration of the global nitrogen cycle: Sources and
consequences. Ecol Appl 7: 737-750. http://dx.doi.org/10.189Q/1051-
0761(1997N)007r0737:HAOTGN12.0.CQ:2.
Vitousek. PM: Howarth. RW. (1991). Nitrogen limitation on land and in the sea: How can it
occur? Biogeochemistry 13: 87-115. http://dx.doi.org/10.1007/BF000Q2772.
Vitousek, PM; Menge, DNL; Reed, SC; Cleveland, CC. (2013). Biological nitrogen fixation:
Rates, patterns and ecological controls in terrestrial ecosystems. Philos Trans R Soc Lond
B Biol Sci 368. http://dx.doi.org/10.1098/rstb.2013.0119.
Vitousek, PM: Porder, S: Houlton, BZ; Chadwick, OA. (2010). Terrestrial phosphorus
limitation: mechanisms, implications, and nitrogenphosphorus interactions. Ecol Appl 20:
5-15. http://dx.doi.Org/10.1890/08-0127.l.
Vitt, DH; Wieder, K; Halsev, LA: Turetsky, M. (2003). Response of Sphagnum fuscum to
nitrogen deposition: A case study of ombrogenous peatlands in Alberta, Canada.
Bryologist 106: 235-245. http://dx.doi.org/10.1639/00Q7-
2745(2003N)106r0235:ROSFTN12.0.CQ;2.
Vivanco, L; Irvine, IC: Martiny, JBH. (2015). Nonlinear responses in salt marsh functioning to
increased nitrogen addition. Ecology 96: 936-947. http://dx.doi.org/10.189Q/13-
1983.l.sm.
Vivas, A: Barea, JM; Biro, B; Azcon, R. (2006). Effectiveness of autochthonous bacterium and
mycorrhizal fungus on Trifolium growth, symbiotic development and soil enzymatic
activities in Zn contaminated soil. J Appl Microbiol 100: 587-598.
http://dx.doi.org/10.1111/j. 1365-2672.2005.02804.x.
Vivas, A: Voros, A: Biro, B; Barea, JM: Ruiz-Lozano, JM: Azcon, R. (2003). Beneficial effects
of indigenous Cd-tolerant and Cd-sensitive Glomus mosseae associated with a Cd-
adapted strain of Brevibacillus sp in improving plant tolerance to Cd contamination. Appl
Soil Ecol 24: 177-186. http://dx.doi.org/10.1016/S0929-1393(03)00088-X.
Vives, I; Grimalt, JO; Ventura, M; Catalan, J. (2005). Distribution of polycyclic aromatic
hydrocarbons in the food web of a high mountain lake, Pyrenees, Catalonia, Spain.
Environ Toxicol Chem 24: 1344-1352. http://dx.doi.Org/10.1897/04-181R.l.
Vlieger, L; Candolin, U. (2009). How not to be seen: Does eutrophication influence three-spined
stickleback Gasterosteus aculeatus sneaking behaviour? J Fish Biol 75: 2163-2174.
http://dx.doi.org/10.1111/j. 1095-8649.2009.02403 ,x.
June 2018
R-230
DRAFT: Do Not Cite or Quote

-------
Vogel-Mikus. K; Drobne. D; Regvar. M. (2005). Zn, Cd and Pb accumulation and arbuscular
mycorrhizal colonisation of pennycress Thlaspi praecox Wulf. (Brassicaceae) from the
vicinity of a lead mine and smelter in Slovenia. Environ Pollut 133: 233-242.
http://dx.doi.Org/10.1016/i.envpol.2004.06.021.
Vogelmikus. K; Pongrac. P; Kump. P; Necemer. M; Regvar. M. (2006). Colonisation of a Zn, Cd
and Pb hyperaccumulator Wulfen with indigenous arbuscular mycorrhizal fungal mixture
induces changes in heavy metal and nutrient uptake. Environ Pollut 139: 362-371.
http://dx.doi.Org/10.1016/i.envpol.2005.05.005.
Yolk. M; Obrist D: Novak. K; Giger. R; Bassin. S: Fuhrer. J. (2011). Subalpine grassland
carbon dioxide fluxes indicate substantial carbon losses under increased nitrogen
deposition, but not at elevated ozone concentration. Global Change Biol 17: 366-376.
http://dx.doi.org/10.111 l/i.l365-2486.2010.02228.x.
Yolk. M; Wolff. V: Bassin. S: Ammann. C: Fuhrer. J. (2014). High tolerance of subalpine
grassland to long-term ozone exposure is independent of N input and climatic drivers.
Environ Pollut 189: 161-168. http://dx.doi.Org/10.1016/i.envpol.2014.02.032.
Volkenborn. N: Polerecky. L; Wethev. D; Dewitt. T; Woodin. S. (2012). Hydraulic activities by
ghost shrimp Neotrypaea californiensis induce oxicanoxic oscillations in sediments. Mar
Ecol Prog Ser455: 141-156. http://dx.doi.org/10.3354/meps09645.
Vollenweider. RA. (1968). Scientific fundamentals of the eutrophication of lakes and flowing
waters, with particular reference to nitrogen and phosphorus as factors in eutrophi cation
(pp. 193.). (DAS/CSI/62.27). Paris, France: Organisation for Economic Co-operation and
Development.
von Ballmoos. P; Nussbaum. S: Brunold. C. (1993). The relationship of nitrate reductase activity
to uptake and assimilation of atmospheric 15N02-nitrogen in needles of Norway Spruce
(Picea Abies [L.] Karst.). Isotopes Environ Health Stud 29: 59-70.
http://dx.doi.org/10.1080/10256019308Q46136.
von Bobrutzki. K; Braban. CF; Famulari. D; Jones. SK; Blackall. T; Smith. TEL: Blom. M; Coe.
H; Gallagher. M; Ghalaieny. M; McGillen. MR: Percival. CJ: Whitehead. JD; Ellis. R;
Murphy. J: Mohacsi. A: Pogany. A: Junninen. H; Rantanen. S: Sutton. MA: Nemitz. E.
(2010). Field inter-comparison of eleven atmospheric ammonia measurement techniques.
Atmos Meas Tech 3: 91-112. http://dx.doi.org/10.5194/amt-3-91-2010.
Von Ruckert. G: Giani. A. (2004). Effect of nitrate and ammonium on the growth and protein
concentration of Microcystis viridis Lemmermann (Cyanobacteria). Revista Brasileira de
Botanica 27: 325-331. http://dx.doi.org/10.1590/S0100-8404200400020Q011.
Vorosmartv. CJ: Leveque. C: Revenga. C. (2005). Chapter 7: Fresh water. In A Millennium
Ecosystem (Ed.), Ecosystems and Human Well-being: Current State and Trends:
Findings of the Condition and Trends Working Group of the Millennium Ecosystem
Assessment (pp. 165-207). Washington, DC: Island Press.
http://www.unep.org/maweb/documents/document.276.aspx.pdf.
Yose. JM; Swank. WT. (1990). Preliminary estimates of foliar absorption of 15N-labeled nitric
acid vapor (HN03) by mature eastern white pine (Pinusstrobus). Can J For Res 20: 857-
860. http://dx.doi.org/10.1139/x90-l 14.
June 2018
R-231
DRAFT: Do Not Cite or Quote

-------
Vourlitis. GL. (2012). Aboveground net primary production response of semi-arid shrublands to
chronic experimental dry-season N input. Ecosphere 3: 1-9.
http://dx.doi.Org/10.1890/ESll-00339.l.
Vourlitis. GL. (2017). Chronic N enrichment and drought alter plant cover and community
composition in a Mediterranean-type semi-arid shrubland. Oecologia 184: 267-277.
http://dx.doi. org/10.1007/s00442-017-3 860-1.
Vourlitis. GL: Fernandez. JS. (2012). Changes in the soil, litter, and vegetation nitrogen and
carbon concentrations of semiarid shrublands in response to chronic dry season nitrogen
input. J Arid Environ 82: 115-122. http://dx.doi.Org/10.1016/i.iaridenv.2012.02.006.
Vourlitis. GL: Fernandez. JS. (2015). Carbon and nitrogen mineralization of semi-arid shrubland
soils exposed to chronic nitrogen inputs and pulses of labile carbon and nitrogen. J Arid
Environ 122: 37-45. http://dx.doi.Org/10.1016/i.iaridenv.2015.06.007.
Vourlitis, GL; Pasquini, SC. (2008). Carbon and nitrogen dynamics of pre- and post-fire
chaparral exposed to varying atmospheric N deposition. J Arid Environ 72: 1448-1463.
http://dx.doi.Org/10.1016/i.iaridenv.2008.02.020.
Vourlitis. GL: Pasquini. SC. (2009). Experimental dry-season N deposition alters species
composition in southern Californian mediterranean-type shrublands. Ecology 90: 2183-
2189. http://dx.doi.Org/10.1890/08-1121.l.
Vourlitis. GL: Pasquini. SC: Mustard. R. (2009). Effects of dry-season N input on the
productivity andN storage of Mediterranean-type shrublands. Ecosystems 12: 473-488.
http://dx.doi.org/10.1007/slQ021-009-9236-6.
Vourlitis. GL: Zorba. G. (2007). Nitrogen and carbon mineralization of semi-arid shrubland soil
exposed to long-term atmospheric nitrogen deposition. Biol Fertil Soils 43: 611-615.
http://dx.doi. org/10.1007/s003 74-006-013 7-v.
Vrba. J: Boikova. J: Chvoika. P; Fott. J. an: Kopacek. J: Macek. M; Nedbalova. L; Papacek. M;
Radkova. V: Sacherova. V: Sol dan. T; Sorf. M. (2016). Constraints on the biological
recovery of the Bohemian Forest lakes from acid stress. Freshw Biol 61: 376-395.
http://dx.doi.org/10. Ill 1/fwb. 12714.
Wakelev. JS. (2002). Developing a "regionalized" version of the Corps of Engineers wetlands
delineation manual: Issues and recommendations. (ERDC/EL TR-02-20). Vicksburg,
MS: U.S. Army Engineer Research and Development Center.
http://www.usace.armv.mi1/Portals/2/docs/civilworks/regulatory/reg supp/dev reg wetla
nds.pdf.
Walker. JT; Jones. MR: Bash. JO: Myles. L; Meyers. T; Schwede. D; Herrick. J: Nemitz. E;
Robarge. W. (2013). Processes of ammonia air-surface exchange in a fertilized Zea mays
canopy. Biogeosciences 10: 981-998. http://dx.doi.org/10.5194/bg-10-981 -2013.
Walker. WE: Harremoes. P; Rotmans. J: Van Per Sluiis. JP: Van Asselt. MBA: Janssen. P;
Kraver Von Krauss. MP. (2003). Defining uncertainty: a conceptual basis for uncertainty
management in model-based decision support. Integrated Assessment 4: 5-17.
http://dx.doi.Org/10.1076/iaii.4.l.5.16466.
June 2018
R-232
DRAFT: Do Not Cite or Quote

-------
Wall CC; Gobler. CJ; Peterson. BJ; Ward. JE. (2013). Contrasting growth patterns of
suspension-feeding molluscs (Mercenaria mercenaria, Crassostrea virginica, Argopecten
irradians, and Crepidula fornicata) across a eutrophication gradient in the Peconic
estuary, NY, USA. Estuaries Coasts 36: 1274-1291. http://dx.doi.org/10.10Q7/sl2237-
013-9632-1.
Wallace. RB; Baumann. H; Grear. JS: Aller. RC: Gobler. CJ. (2014). Coastal ocean acidification:
The other eutrophi cation problem. Estuar Coast Shelf Sci 148: 1-13.
http://dx.doi.Org/10.1016/i.ecss.2014.05.027.
Wallace. ZP; Lovett. GM; Hart. JE: Machona. B. (2007). Effects of nitrogen saturation on tree
growth and death in a mixed-oak forest. For Ecol Manage 243: 210-218.
http://dx.doi.Org/10.1016/i.foreco.2007.02.015.
Wallenda. T; Kottke. I. (1998). Nitrogen deposition and ectomycorrhizas. NewPhytol 139: 169-
187. http://dx.doi.Org/10.1046/i.1469-8137.1998.00176.x.
Waller. K; Driscoll. C: Lynch. J: Newcomb. D; Roy. K. (2012). Long-term recovery of lakes in
the Adirondack region of New York to decreases in acidic deposition. Atmos Environ 46:
56-64. http://dx.doi.Org/10.1016/i.atmosenv.2011.10.031.
Wallman, P; Svensson, MGE; Sverdrup, H; Belyazid, S. (2005). ForSAFE - an integrated
process-oriented forest model for long-term sustainability assessments. For Ecol Manage
207: 19-36. http://dx.doi.Org/10.1016/i.foreco.2004.10.016.
Walter. CA; Raiff. DT; Burnham. MB: Gilliam. FS: Adams. MB: Peteriohn. WT. (2016).
Nitrogen fertilization interacts with light to increase Rubus spp. cover in a temperate
forest. Plant Ecol 217: 421-430. http://dx.doi.org/10.1007/sll258-016-Q583-z.
Wamelink. GWW: Goedhart. PW: Malinowska. AH: Frissel. JY; Wegman. RJM; Slim. PA: van
Dobben. HF. (2011). Ecological ranges for the pH and NO(3) of syntaxa: A new basis for
the estimation of critical loads for acid and nitrogen deposition. J Veg Sci 22: 741-749.
http://dx.doi.org/10.1111/i. 1654-1103.2011.01286.x.
Wamelink. GWW: van Dobben. HF: Mol-Dijkstra. JP; Schouwenberg. E; Kros. J: de Vries. W:
Berendse. F. (2009a). Effect of nitrogen deposition reduction on biodiversity and carbon
sequestration. For Ecol Manage 258: 1774-1779.
http://dx.doi.Org/10.1016/i.foreco.2008.10.024.
Wamelink. GWW: Wieggers. HJJ; Reinds. GJ: Kros. J: Mol-Dijkstra. JP: van Oiien. M; de
Vries. W. (2009b). Modelling impacts of changes in carbon dioxide concentration,
climate and nitrogen deposition on carbon sequestration by European forests and forest
soils. For Ecol Manage 258: 1794-1805. http://dx.doi.Org/10.1016/i.foreco.2009.05.018.
Wang. C: Lei. A: Zhou. K: Hu. Z: Hao. W: Yang. J. (2014a). Growth and nitrogen uptake
characteristics reveal outbreak mechanism of the opportunistic macroalga Gracilaria
tenuistipitata. PLoS ONE 9: el08980. http://dx.doi.org/10.1371/iournal.pone.0108980.
Wang. D: He. HL: Gao. O: Zhao. CZ: Zhao. WO: Yin. CY: Chen. XL: Ma. ZL: Li. DP: Sun.
DP: Cheng. XY: Liu. O. (2017a). Effects of short-term N addition on plant biomass
allocation and C and N pools of the Sibiraea angustata scrub ecosystem. Eur J Soil Sci
68: 212-220. http://dx.doi.org/10.1111/eiss. 12414.
June 2018
R-233
DRAFT: Do Not Cite or Quote

-------
Wang. G; Xue. S. ha; Liu. F; Liu. G. (2017b). Nitrogen addition increases the production and
turnover of the lower-order roots but not of the higher-order roots of Bothriochloa
ischaemum. Plant Soil 415: 423-434. http://dx.doi.org/10.10Q7/sl 1104-016-3160-2.
Wang. H; Yu. L; Zhang. Z; Liu. W: Chen. L; Cao. G: Yue. H; Zhou. J: Yang. Y; Tang. Y; He.
JS. (2017c). Molecular mechanisms of water table lowering and nitrogen deposition in
affecting greenhouse gas emissions from a Tibetan alpine wetland. Global Change Biol
23: 815-829. http://dx.doi.Org/10.l 111/gcb. 13467.
Wang. J: Solomon. D; Lehmann. J: Zhang. X: Amelung. W. (2006). Soil organic sulfur forms
and dynamics in the Great Plains of North America as influenced by long-term
cultivation and climate. Geoderma 133: 160-172.
http://dx.doi.Org/10.1016/i.geoderma.2005.07.003.
Wang. M; Larmola. T; Murphy. MT; Moore. T. imR; Bubier. JL. (2016a). Stoichiometric
response of shrubs and mosses to long-term nutrient (N, P and K) addition in an
ombrotrophic peatland. Plant Soil 400: 403-416. http://dx.doi.org/10.1007/sl 1104-015-
2744-6.
Wang. M; Murphy. MT: Moore. TR. (2014b). Nutrient resorption of two evergreen shrubs in
response to long-term fertilization in a bog. Oecologia 174: 365-377.
http://dx.doi. org/10.1007/s00442-013-2784-7.
Wang. M; Shi. S: Lin. F. ei; Jiang. P. (2014c). Response of the soil fungal community to multi-
factor environmental changes in a temperate forest. Appl Soil Ecol 81: 45-56.
http://dx.doi.Org/10.1016/i.apsoil.2014.04.008.
Wang. R; Creamer. CA; Wang. X. ue: He. P; Xu. Z; Jiang. Y. (2016b). The effects of a 9-year
nitrogen and water addition on soil aggregate phosphorus and sulfur availability in a
semi-arid grassland. Ecol Indicat 61: 806-814.
http://dx.doi.Org/10.1016/i.ecolind.2015.10.033.
Wang. R; Dorodnikov. M; Dijkstra. FA: Yang. S: Xu. Z; Li. H. ui; Jiang. Y. (2017d).
Sensitivities to nitrogen and water addition vary among microbial groups within soil
aggregates in a semiarid grassland. Biol Fertil Soils 53: 129-140.
http://dx.doi.org/10.1007/sQ0374-016-1165-x.
Wang. Y. -S: Cheng. S. -L: Yu. G. -R: Fang. H. -J: Mo. J. -M: Xu. M. -J: Gao. W. -L. (2015a).
Response of carbon utilization and enzymatic activities to nitrogen deposition in three
forests of subtropical China. Can J For Res 45: 394-401. http://dx.doi.org/10.1139/cifr-
2014-0445.
Wang. Y; Jiang. O; Yang. Z; Sun. W: Wang. D. (2015b). Effects of water and nitrogen addition
on ecosystem carbon exchange in a meadow steppe. PLoS ONE 10: e0127695.
http://dx.doi.org/10.1371/iournal.pone.0127695.
Warby, RAF; Johnson, CE; Driscoll, CT. (2008). Changes in aluminum concentrations and
speciation in lakes across the Northeastern US following reductions in acidic deposition.
Environ Sci Technol 42: 8668-8674. http://dx.doi.org/10.1021/es801125d.
Warby. RAF: Johnson. CE: Driscoll. CT. (2009). Continuing acidification of organic soils across
the northeastern USA: 1984-2001. Soil Sci Soc Am J 73: 274-284.
http://dx.doi.org/10.2136/sssai2007.0Q16.
June 2018
R-234
DRAFT: Do Not Cite or Quote

-------
Ward. BB. (2013). How nitrogen is lost. Science 341: 352-353.
http://dx.doi.org/10.1126/science. 1240314.
Wardle, DA: Gundale, MJ; Jaderlund, A: Nilsson, M, -C. (2013). Decoupled long-term effects of
nutrient enrichment on aboveground and belowground properties in subalpine tundra.
Ecology 94: 904-919. http://dx.doi.Org/10.1890/12-0948.l.
Wardle. DA: Jonsson. M; Mayor. JR; Metcalfe. DB. (2016). Above-ground and below-ground
responses to long-term nutrient addition across a retrogressive chronosequence. J Ecol
104: 545-560. http://dx.doi.Org/10.l 111/1365-2745.12520.
Warneck, P. (1999). The relative importance of various pathways for the oxidation of sulfur
dioxide and nitrogen dioxide in sunlit continental fair weather clouds. Phys Chem Chem
Phys 1: 5471-5483. http://dx.doi.org/10.1039/a906558i.
Warner. KA; Bonzongo. JCJ; Roden. EE: Ward. GM; Green. AC: Chaubev. I: Lyons. WB;
Arlington, DA. (2005). Effect of watershed parameters on mercury distribution in
different environmental compartments in the Mobile Alabama River Basin, USA. Sci
Total Environ 347: 187-207. http://dx.doi.Org/10.1016/i.scitotenv.2004.12.011.
Warren. PR: Kraft. CE; Josephson. DC: Driscoll. CT. (2017). Acid rain recovery may help to
mitigate the impacts of climate change on thermally sensitive fish in lakes across eastern
North America. Global Change Biol 23: 2149-2153. http://dx.doi.Org/10.l 111/gcb. 13568.
Warren. PR: Likens. GE; Buso. DC: Kraft. CE. (2008). Status and distribution of fish in an acid-
impacted watershed of the northeastern United States (Hubbard Brook, NH). Northeast
Nat 15: 375-390. http://dx.doi.Org/10.1656/1092-6194-15.3.375.
Warren. PR: Mineau. MM: Ward. EJ: Kraft. CE. (2010). Relating fish biomass to habitat and
chemistry in headwater streams of the northeastern United States. Environmental Biology
of Fishes 88: 51-62. http://dx.doi.org/10.1007/slQ641-010-9617-x.
Warry. FY: Reich. P: Cook. PLM: Mac Nallv. R: Thomson. JR: Woodland. RJ. (2016). Nitrogen
loads influence trophic organization of estuarine fish assemblages. Funct Ecol 30: 1723-
1733. http://dx.doi.org/10.1111/1365-2435.12647.
Wasik. JC: Engstrom. PR: Mitchell. CPJ: Swain. EB: Monson. BA: Balogh. SJ: Jeremiason. JP:
Branfireun. BA: Kolka. RK: Almendinger. JE. (2015). The effects of hydrologic
fluctuation and sulfate regeneration on mercury cycling in an experimental peatland. Jour
Geo Res: Biog 120: 16971715. http://dx.doi.org/10.1002/2015JG0Q2993.
Wasik. JKC: Mitchell. CPJ: Engstrom. PR: Swain. EB: Monson. BA: Balogh. SJ: Jeremiason.
JP: Branfireun. BA: Eggert. SL: Kolka. RK: Almendinger. JE. (2012). Methylmercury
declines in a boreal peatland when experimental sulfate deposition decreases. Environ Sci
Technol 46: 6663-6671. http://dx.doi.org/10.1021/es300865f.
Wason. J. avW: Povciak. M: Beier. CM: Battles. JJ. (2017). Tree growth is more sensitive than
species distributions to recent changes in climate and acidic deposition in the northeastern
United States. J Appl Ecol 54: 1648-1657. http://dx.doi.org/10.1111/1365-2664.12899.
Watmough, SA: Meadows, MJ. (2014). Po earthworms have a greater influence on nitrogen
dynamics than atmospheric nitrogen deposition? Ecosystems 17: 1257-1270.
http://dx.doi. org/10.1007/s 10021 -014-9794-0.
June 2018
R-235
DRAFT: Do Not Cite or Quote

-------
Watmough. SA; Whitfield. CJ; Fenn. ME. (2014). The importance of atmospheric base cation
deposition for preventing soil acidification in the Athabasca Oil Sands Region of Canada.
Sci Total Environ 493: 1-11. http://dx.doi.ore/10.1016/i.scitotenv.2014.05.110.
Watras. CJ; Bloom. NS. (1992). Mercury and methylmercury in individual zooplankton:
Implications for bioaccumulation. Limnol Oceanogr37: 1313-1318.
http://dx.doi.Org/10.4319/lo.1992.37.6.1313.
Watras. CJ: Bloom. NS: Claas. SA: Morrison. KA; Gilmour. CC: Craig. SR. (1995).
Methylmercury production in the anoxic hypolimnion of a dimictic seepage lake. Water
Air Soil Pollut 80: 735745. http://dx.doi.org/10.1007/BF01189725.
Watras. CJ; Morrison. KA. (2008). The response of two remote, temperate lakes to changes in
atmospheric mercury deposition, sulfate, and the water cycle. Can J Fish Aquat Sci 65:
100-116. http://dx.doi.org/10.1139/F07-159.
Watras, CJ; Morrison, KA; Regnell, O; Kratz, TK. (2006). The methylmercury cycle in Little
Rock Lake during experimental acidification and recovery. Limnology 51: 257-270.
http://dx.doi.Org/10.4319/lo.2006.51.l.0257.
Wavcott, M; Duarte, CM; Carruthers, TJ; Orth, RJ; Dennison, WC; Olyarnik, S; Calladine, A;
Fourqurean, JW; Heck, KL, Jr; Hughes, AR; Kendrick, GA; Kenworthy, WJ; Short, FT;
Williams, SL. (2009). Accelerating loss of seagrasses across the globe threatens coastal
ecosystems. ProcNatl Acad Sci USA 106: 12377-12381.
http://dx.doi.org/10.1073/pnas.09056201Q6.
Weand, MP; Arthur, MA; Lovett, GM; McCulley, RL; Weathers, KC. (2010). Effects of tree
species and N additions on forest floor microbial communities and extracellular enzyme
activities. Soil Biol Biochem 42: 2161-2173.
http://dx.doi.Org/10.1016/i.soilbio.2010.08.012.
Weathers, KC; Simkin, SM; Lovett, GM; Lindberg, SE. (2006). Empirical modeling of
atmospheric deposition in mountainous landscapes. Ecol Appl 16: 1590-1607.
http://dx.doi.org/10.1890/1051-0761(2006N)016ri590:EMOADI12.0.CQ:2.
Webb, JR; Cosby, BJ; Devinev, FA, Jr; Galloway, JN; Maben, SW; Bulger, AJ. (2004). Are
brook trout streams in western Virginia and Shenandoah National Park recovering from
acidification? Environ Sci Technol 38: 4091-4096. http://dx.doi.org/10.1021/es049958a.
Weese, DJ; Heath, KD; Dentinger, BTM; Lau, JA. (2015). Long-term nitrogen addition causes
the evolution of less-cooperative mutualists. Evolution 69: 631-642.
http://dx.doi.org/10. Ill 1/evo. 12594.
Wei. C; Yu. O; Bai. E; Lu. X; Li. O; Xia. J; Kardol. P; Liang. W; Wang. Z; Han. X. (2013).
Nitrogen deposition weakens plant-microbe interactions in grassland ecosystems. Global
Change Biol 19: 3688-3697. http://dx.doi.org/10. Ill 1/gcb. 12348.
Weier. KL; Doran. JW; Power. JF; Walters. DT. (1993). Denitrification and the
dinitrogen/nitrous oxide ratio as affected by soil water, available carbon, and nitrate. Soil
Sci Soc Am J 57: 66-72. http://dx.doi.org/10.2136/sssail993.03615995005700010Q13x.
Weinstock. L. (2012). Putting multipollutant monitoring into practice: EPA's NCore network.
EM Magazine May 2012: 26-31.
June 2018
R-236
DRAFT: Do Not Cite or Quote

-------
Weiss. SB. (1999). Cars, cows, and checkerspot butterflies: Nitrogen deposition and
management of nutrient-poor grasslands for a threatened species. Conserv Biol 13: 1476-
1486. http://dx.doi.Org/10.1046/i.1523-1739.1999.98468.x.
Wellburn. AR. (1990). Tansley review no 24: Why are atmospheric oxides of nitrogen usually
phytotoxic and not alternative fertilizers? [Review], New Phytol 115: 395-429.
http://dx.doi.Org/10.llll/i.1469-8137.1990.tb00467.x.
Welter. JR: Benstead. JP; Cross. WF; Hood. JM; Huryn. AD: Johnson. PW: Williamson. TJ.
(2015). Does N2 fixation amplify the temperature dependence of ecosystem metabolism?
Ecology 96: 603-610. http://dx.doi.Org/10.1890/14-1667.l.
Welti. N: Haves. M; Lockington. D. (2017). Seasonal nitrous oxide and methane emissions
across a subtropical estuarine salinity gradient. Biogeochemistry 132: 55-69.
http://dx.doi.org/10.1007/slQ533-016-0287-4.
Welz, O; Savee, JD; Osborn, PL; Vasu, SS; Percival, CJ; Shallcross, DE; Taatjes, CA. (2012).
Direct kinetic measurements of Criegee intermediate (CHOO) formed by reaction of CHI
with O. Science 335: 204-207. http://dx.doi.Org/10.l 126/science. 1213229.
Wendel. S: Moore. T; Bubier. J: Blodau. C. (2011). Experimental nitrogen, phosphorus, and
potassium deposition decreases summer soil temperatures, water contents, and soil C02
concentrations in a northern bog. Biogeosciences 8: 585-595.
http://dx.doi.org/10.5194/bg-8-585-2011.
Wentz. DA: Brigham. ME: Chasar. LC: Lutz. MA: Krabbenhoft. DP. (2014). Mercury in the
nations streamsLevels, trends, and implications. (Circular 1935). Reston, VA: U.S.
Geological Survey, http://dx.doi.org/10.3133/cirl395.
Werth. S: Sork. VL. (2008). Local genetic structure in a North American epiphytic lichen,
Ramalina menziesii (Ramalinaceae). Am J Bot 95: 568-576.
http://dx.doi.org/10.3732/aib.2007Q24.
Westerling. AL; Hildalgo. HG: Cavan. PR: Swetnam. TW. (2006). Warming and earlier spring
increase western US forest wildfire activity. Science 313: 940-943.
http://dx.doi.org/10.1126/science. 1128834.
Wetz. MS: Yoskowitz. DW. (2013). An 'extreme' future for estuaries? Effects of extreme
climatic events on estuarine water quality and ecology [Review], Mar Pollut Bull 69: 7-
18. http://dx.doi.Org/10.1016/i.marpolbul.2013.01.020.
Wevens. N: Thiis. S: Popek. R; Witters. N: Przybvsz. A: Espenshade. J: Gawronska. H;
Vangronsveld. J: Gawronski. SW. (2015). The Role of Plant-Microbe Interactions and
Their Exploitation for Phytoremediation of Air Pollutants [Review], International Journal
of Molecular Sciences 16: 25576-25604. http://dx.doi.org/10.3390/iimsl61025576.
Wevhenmever. GA; Karlsson. J. an. (2009). Nonlinear response of dissolved organic carbon
concentrations in boreal lakes to increasing temperatures. Limnol Oceanogr 54: 2513-
2519.
June 2018
R-237
DRAFT: Do Not Cite or Quote

-------
Whigham, DF; Verhoeven. JTA; Samarkin. V; Megonigal, PJ. (2009). Responses of Avicennia
germinans (Black Mangrove) and the soil microbial community to nitrogen addition in a
hypersaline wetland. Estuaries Coasts 32: 926-936. http://dx.doi.org/10.10Q7/sl2237-
009-9184-6.
Whitall, D; Castro. M; Driscoll, C. (2004). Evaluation of management strategies for reducing
nitrogen loadings to four US estuaries. Sci Total Environ 333: 25-36.
http://dx.doi.Org/10.1016/i.scitotenv.2004.05.006.
White. MJ: Santhi. C: Kantian. N: Arnold. JG: Harm el. D; Norfleet L; Allen. P; Diluzio. M;
Wang. X: Atwood. J: Hanev. E; Johnson. MY. (2014). Nutrient delivery from the
Mississippi River to the Gulf of Mexico and effects of cropland conservation. J Soil
Water Conserv 69: 26-40. http://dx.doi.Org/10.2489/iswc.69.l.26.
Whitehead. PG: Wilbv. RL; Battarbee. RW: Kernan. M; Wade. AJ. (2009). A review of the
potential impacts of climate change on surface water quality. Hydrological Sciences
Journal 54: 101-123. http://dx.doi.Org/10.1623/hysi.54.l.101.
Whitfield. C. ,J.. P..helan. J...N.. B..ucklev. J.... C. L..ark. C...M.. G..uthrie. S.... L..ynch. J...A.
(2018). Estimating base cation weathering rates in the USA: challenges of uncertain soil
mineralogy and specific surface area with applications of the profile model. Water Air
Soil Pollut 229: 61. http://dx.doi.Org/https://doi.org/10.1007/sll270-018-3691-7.
Whitfield. CJ: Aherne. J: Cosby. BJ; Watmough. SA. (2010a). Modelling catchment response to
acid deposition: A regional dual application of the MAGIC model to soils and lakes in
the Athabasca Oil Sands Region, Alberta. Journal of Limnology 69: 147-160.
http://dx.doi.org/10.4081/ilimnol.2010.sl.147.
Whitfield. CJ: Aherne. J: Watmough. SA. (2009). Modeling soil acidification in the Athabasca
Oil Sands region, Alberta, Canada. Environ Sci Technol 43: 5844-5850.
http://dx.doi.org/10.1021/es9005652.
Whitfield. CJ: Aherne. J: Watmough. SA: Mcdonald. M. (2010b). Estimating the sensitivity of
forest soils to acid deposition in the Athabasca Oil Sands Region, Alberta. Journal of
Limnology 69: 201-208. http://dx.doi.org/10.3274/JL10-69-Sl-20.
Whitfield, CJ: Watmough, SA. (2012). A regional approach for mineral soil weathering
estimation and critical load assessment in boreal Saskatchewan, Canada. Sci Total
Environ 437: 165-172. http://dx.doi.Org/10.1016/i.scitotenv.2012.08.046.
Whitmore, ME: Mansfield, TA. (1983). Effects of long-term exposures to S02 and N02 on Poa
pratensis and other grasses. Environ Pollut 31: 217-235. http://dx.doi.org/10.1016/Q143-
1471(83^)90078-8.
Whittinghill, KA; Currie, WS: Zak, PR: Burton, AJ: Pregitzer, KS. (2012). Anthropogenic N
deposition increases soil C storage by decreasing the extent of litter decay: Analysis of
field observations with an ecosystem model. Ecosystems 15: 450-461.
http://dx.doi.org/10.1007/slQ021-012-9521-7.
WHO (World Health Organization). (1998). Guidelines for drinking-water quality: Second
edition: Addendum to volume 2: Health criteria and other supporting information.
(WHO/EOS/98.1). Geneva, Switzerland, http://apps.who.int/iris/handle/10665/63844.
June 2018
R-238
DRAFT: Do Not Cite or Quote

-------
Wiedenmann. J; P'Angelo, C; Smith. EG; Hunt AN; Legiret FE; Postle. AD; Achterberg, EP.
(2013). Nutrient enrichment can increase the susceptibility of reef corals to bleaching.
Nat Clim Chang 3: 160-164. http://dx.doi.org/10.1038/NCLIMATE1661.
Wieder. RK; Vile. MA; Scott. KD; Albright. CM; Mcmillen. K; Vitt. DH; Fenn. M. (2016).
Differential Effects of High Atmospheric N and S Deposition on Bog Plant/Lichen Tissue
and Porewater Chemistry Across the Athabasca Oil Sands Region. Environ Sci Technol
50: 12630-12640. http://dx.doi.org/10.1021/acs.est.6b03109.
Wiener. JG; Fitzgerald. WF; Watras. CJ; Rada. RG. (1990). Partitioning and bioavailability of
mercury in an experimentally acidified Wisconsin lake. Environ Toxicol Chem 9: 909-
918. http://dx.doi.org/10.1002/etc.56200907Q9.
Wigand. C. (2008). Coastal salt marsh community change in Narragansett Bay in response to
cultural eutrophication. In SPRINGER SERIES ON ENVIRONMENTAL
MANAGEMENT.
Wigand. C; Brennan. P; Stolt. M; Holt. M; Ryba. S. (2009). Soil respiration rates in coastal
marshes subject to increasing watershed nitrogen loads in Southern New England, USA.
Wetlands 29: 952-963. http://dx.doi.Org/10.1672/08-147.l.
Wigand, C; Davey, E; Johnson, R; Sundberg, K; Morris, J; Kenny, P; Smith, E; Holt, M. (2015).
Nutrient effects on belowground organic matter in a minerogenic salt marsh, North Inlet,
SC. Estuaries Coasts 38: 1838-1853. http://dx.doi.org/10.1007/sl2237-014-9937-8.
Wigand, C; McKinnev, R; Chintala, M; Lussier, S; Heltshe, J. (2010). Development of a
reference coastal wetland set in Southern New England (USA). Environ Monit Assess
161: 583-598. http://dx.doi.org/10.1007/s 10661-009-0770-7.
Wigand, C; McKinnev, RA; Charpentier, MA; Chintala, MM; Thursbv, GB. (2003).
Relationships of nitrogen loadings, residential development, and physical characteristics
with plant structure in new England salt marshes. Estuaries 26: 1494-1504.
http://dx.doi.org/10.1007/BF028Q3658.
Wigington, PJ; Baker, JP; DeWalle, PR; Kretser, WA; Murdoch, PS; Siminon, HA; Van Sickle,
J; McDowell, MK; Peck, DV; Barchet, WR. (1993). Episodic acidification of streams in
the northeastern United States: Chemical and biological results of the Episodic Response
Project. (EPA/600/R-93/190). Washington, DC: U.S. Environmental Protection Agency.
Wigington, PJ, Jr; Baker, JP; DeWalle, PR; Kretser, WA; Murdoch, PS; Simonin, HA; Van
Sickle, J; McDowell, MK; Peck, DV; Barchet, WR. (1996a). Episodic acidification of
small streams in the northeastern United States: Episodic Response Project. Ecol Appl 6:
374-388. http://dx.doi.org/10.2307/2269377.
Wigington, PJ, Jr; Pavies, TP; Tranter, M; Eshleman, KN. (1990). Episodic acidification of
surface waters due to acidic deposition. In State of Science/Technology Report 12.
(NAPAP REPORT 12). Washington, PC: National Acid Precipitation Assessment
Program.
Wigington, PJ, Jr; PeWalle, PR; Murdoch, PS; Kretser, WA; Simonin, HA; Van Sickle, J;
Baker, JP. (1996b). Episodic acidification of small streams in the northeastern United
States: Ionic controls of episodes. Ecol Appl 6: 389-407.
http://dx.doi.org/10.2307/2269378.
June 2018
R-239
DRAFT: Do Not Cite or Quote

-------
Wilkerson. FP; Dugdale, RC; Parker. AE; Blaser. SB; Pimenta. A. (2015). Nutrient uptake and
primary productivity in an urban estuary: Using rate measurements to evaluate
phytoplankton response to different hydrological and nutrient conditions. Aquatic
Ecology 49: 211-233. http://dx.doi.org/10.1007/slQ452-015-9516-5.
Wilkins, K; Aherne, J. (2016). Vegetation community change in Atlantic oak woodlands along a
nitrogen deposition gradient. Environ Pollut 216: 115-124.
http://dx.doi.Org/10.1016/i.envpol.2016.05.024.
Wilkinson. TAC: Wiken. E; Bezaurv-Creel. J: Hourigan, TF; Agardv, T; Herrmann. H;
Janishevski. L; Madden. C: Morgan. L; Padilla. M. (2009). Marine ecoregions of North
America. Montreal, Canada: Commission for Environmental Cooperation.
http://www3.cec.org/islandora/en/item/3256-marine-ecoregions-north-america-en.pdf.
Will-Wolf. S: Jovan. S: Neitlich. P; Peck. JE; Rosentreter. R. (2015). Lichen-based indices to
quantify responses to climate and air pollution across northeastern USA. Bryologist 118:
59-82. http://dx.doi.Org/10.1639/0007-2745-118.l.059.
Willev. JD; Glinski. DA: Southwell. M; Long. MS: Avery. B. Jr; Kieber. RJ. (2011). Decadal
variations of rainwater formic and acetic acid concentrations in Wilmington, NC, USA.
Atmos Environ 45: 1010-1014. http://dx.doi.Org/10.1016/i.atmosenv.2010.10.047.
Williams. JJ; Beutel. M; Nurse. A: Moore. B; Hampton. SE; Saros. JE. (2016a). Phytoplankton
responses to nitrogen enrichment in Pacific Northwest, USA Mountain Lakes.
Hydrobiologia 776: 261-276. http://dx.doi.org/10.1007/slQ750-016-2758-v.
Williams, JJ: Chung, SH; Johansen, AM; Lamb, BK; Vaughan, JK; Beutel, M. (2017a).
Evaluation of atmospheric nitrogen deposition model performance in the context of US
critical load assessments. Atmos Environ 150: 244-255.
http://dx.doi.Org/10.1016/i.atmosenv.2016.l 1.051.
Williams, JJ: Lynch, JA; Saros, JE; Labou, SG. (2017b). Critical loads of atmospheric N
deposition for phytoplankton nutrient limitation shifts in western U.S. mountain lakes.
Ecosphere 8: e01955. http://dx.doi.org/10.1002/ecs2.1955.
Williams, JJ: Nurse, A: Saros, JE: Riedel, J, on: Beutel, M. (2016b). Effects of glaciers on
nutrient concentrations and phytoplankton in lakes within the Northern Cascades
Mountains (USA). Biogeochemistry 131: 373-385. http://dx.doi.org/10.1007/slQ533-016-
0264-v.
Williams, MR: Filoso, S: Longstaff, BJ; Dennison, WC. (2010). Long-term trends of water
quality and biotic metrics in Chesapeake Bay: 1986 to 2008. Estuaries Coasts 33: 1279-
1299. http://dx.doi.org/10.1007/s 12237-010-9333-v.
Williams, MW: Baron, JS: Caine, N; Sommerfeld, R; Sanford, R, Jr. (1996a). Nitrogen
saturation in the Rocky Mountains. Environ Sci Technol 30: 640-646.
http://dx.doi.org/10.1021/es950383e.
Williams, MW; Knauf, M; Cory, R; Caine, N; Liu, F. (2007). Nitrate content and potential
microbial signature of rock glacier outflow, Colorado Front Range. Earth Surface
Processes and Landforms 32: 1032-1047. http://dx.doi.org/10.1002/esp.1455.
June 2018
R-240
DRAFT: Do Not Cite or Quote

-------
Williams. MW; Losleben, M; Caine, N; Greenland. D. (1996b). Changes in climate and
hydrochemical responses in a high-elevation catchment in the Rocky Mountains, USA.
Limnol Oceanogr41: 939-946.
Williams. MW; Platts-Mills. T; Caine. N. (1996c). Landscape controls on surface water nitrate
concentrations at catchment and regional scales in the Colorado Rocky
Mountains.Chapman Conference on "Nitrogen Cycling in Forested Catchments",
September 11-15, 1996, Bend, OR.
Williams. MW; Tonnessen. KA. (2000). Critical loads for inorganic nitrogen deposition in the
Colorado Front Range, USA. Ecol Appl 10: 1648-1665. http://dx.doi.org/10.189Q/1051-
0761(2000^)01 On 648 :CLFIND12.0.CQ;2.
Williams. WD. (2001). Anthropogenic salinisation of inland waters. In Developments in
Hydrobiology: Saline Lakes. Dordrecht, The Netherlands: Springer Netherlands.
http://dx.doi.org/10.1007/978-94-Q17-2934-5 30.
Williamson. TJ; Cross. WF; Benstead. JP; Gislason. GM; Hood. JM; Huryn. AD; Johnson. PW;
Welter. JR. (2016). Warming alters coupled carbon and nutrient cycles in experimental
streams. Global Change Biol 22: 2152-2164. http://dx.doi.org/10. Ill 1/gcb. 13205.
Williston, P; Aherne, J; Watmough, S; Marmorek, D; Hall, A; Bueno, P. delaC; Murray, C;
Henolson, A; Laurence, JA. (2016). Critical levels and loads and the regulation of
industrial emissions in northwest British Columbia, Canada. Atmos Environ 146: 311-
323. http://dx.doi.Org/10.1016/i.atmosenv.2016.08.058.
Wilson, WF; Raymond, PA; Saiers, JE; Sobczak, WV; Xu, N. (2016). Increases in humic and
bioavailable dissolved organic matter in a forested New England headwater stream with
increasing discharge. Mar Freshwat Res 67: 1279-1292.
http://dx.doi.org/10.1071/MF15286.
Wiman, BLB; Agren, GI. (1985). Aerosol depletion and deposition in forests: A model analysis.
Atmos Environ 19: 335-347. http://dx.doi.org/10.1016/0004-6981(85^)90101-5.
Windham-Myers, L; Marvin-Dipasquale, M; Kakouros, E; Agee, JL; Kieu, LH; Strieker, CA;
Fleck, JA; Ackerman, JT. (2014a). Mercury cycling in agricultural and managed
wetlands of California, USA: Seasonal influences of vegetation on mercury methylation,
storage, and transport. Sci Total Environ 484: 308-318.
http://dx.doi.Org/10.1016/i.scitotenv.2013.05.027.
Windham-Myers, L; Marvin-Dipasquale, M; Strieker, CA; Agee, JL; Kieu, LH; Kakouros, E.
(2014b). Mercury cycling in agricultural and managed wetlands of California, USA:
Experimental evidence of vegetation-driven changes in sediment biogeochemistry and
methylmercury production. Sci Total Environ 484: 300-307.
http://dx.doi.Org/10.1016/i.scitotenv.2013.05.028.
Winer, AM; Peters, JW; Smith, JP; Pitts, JN, Jr. (1974). Response of commercial
chemiluminescentNO-N02 analyzers to other nitrogen-containing compounds. Environ
Sci Technol 8: 1118-1121. http://dx.doi.org/10.1021/es60098a0Q4.
June 2018
R-241
DRAFT: Do Not Cite or Quote

-------
Winfield, U; Hatelev, J; Fletcher. JM; James. JB; Bean. CW; Clabburn, P. (2010). Population
trends of Arctic charr (Salvelinus alpinus) in the UK: Assessing the evidence for a
widespread decline in response to climate change. Hydrobiologia 650: 55-65.
http://dx.doi.org/10.1007/sl0750-009-0Q78-l.
Wislicenus. H. (1914). Sammlung von Abhandlungen uber Abgase und Rauschaden. Berlin,
Germany: P Parey.
Wissinger, BP: Eigenbrode, SD; Marshall. JD; Hoines. JD; Newingham, BA. (2014). Altered
nitrogen and precipitation along urban gradients affect harvester ants and seed sources. J
Arid Environ 104: 96-105. http://dx.doi.Org/10.1016/i.iaridenv.2014.02.001.
Wold. AP; Hershev. AE. (1999). Spatial and temporal variability of nutrient limitation in 6 North
Shore tributaries to Lake Superior. J North Am Benthol Soc 18: 2-14.
http://dx.doi.org/10.2307/1468005.
Wolfe, AP; Baron, JS; Cornett, RJ. (2001). Anthropogenic nitrogen deposition induces rapid
ecological changes in alpine lakes of the Colorado Front Range (USA). J Paleolimnol 25:
1-7. http://dx.doi.Org/10.1023/A:1008129509322.
Wolfe, AP: Van Gorpe, AC: Baron, JS. (2003). Recent ecological and biogeochemical changes
in alpine lakes of Rocky Mountain National Park (Colorado, USA): A response to
anthropogenic nitrogen deposition. Geobiology 1: 153-168.
http://dx.doi.Org/10.1046/i.1472-4669.2003.00012.x.
Wolff, GT. (1984). On the nature of nitrate in coarse continental aerosols. Atmos Environ 18:
977-981. http://dx.doi.org/10.1016/0004-6981(84)90073-8.
Wollheim, WM; Vorosmartv, CJ: Bouwman, AF; Green, P; Harrison, J: Linder, E; Peterson, BJ;
Seitzinger, SP; Switski, JPM. (2008). Global N removal by freshwater aquatic systems
using a spatially distributed, within-basin approach. Global Biogeochem Cycles 22.
http://dx.doi.org/10.1029/2007GB0Q2963.
Wong, BBM; Tuomainen, U; Candolin, U. (2012). Algal blooms impact the quality of nest
construction in three-spined sticklebacks. Anim Behav 84: 1541-1545.
http://dx.doi.Org/10.1016/i.anbehav.2012.09.031.
Wood, CM: Al-Reasi, HA: Smith, DS. (2011). The two faces of DOC. Aquat Toxicol 105: 3-8.
http://dx.doi.Org/10.1016/i.aquatox.2011.03.007.
Wood, CM; McDonald, DG. (1987). The physiology of acid/aluminum stress in trout. Ann Soc
RoyZool Bel 117: 399-410.
Wood, JD: Franklin, RB; Garman, G: Mclninch, S: Porter, AJ; Bukaveckas, PA. (2014).
Exposure to the cyanotoxin microcystin arising from interspecific differences in feeding
habits among fish and shellfish in the James River Estuary, Virginia. Environ Sci
Technol 48: 5194-5202. http://dx.doi.org/10.1021/es403491k.
Wood, YA; Meixner, T; Shouse, PJ: Allen, EB. (2006). Altered ecohydrologic response drives
native shrub loss under conditions of elevated nitrogen deposition. J Environ Qual 35: 76-
92.
June 2018
R-242
DRAFT: Do Not Cite or Quote

-------
Woodland. RJ; Thomson. J. R.; Mac Nallv. R; Reich. P; Evrard. V; Wary. FY; Walker. JP;
Cook. PLM. (2015). Nitrogen loads explain primary productivity in estuaries at the
ecosystem scale. Limnol Oceanogr 60: 1751-1762. http://dx.doi.org/10.10Q2/lno.10136.
Woodlev. SK; Freeman. P; Ricciardella. LF. (2014). Environmental acidification is not
associated with altered plasma corticosterone levels in the stream-side salamander,
Desmognathus ochrophaeus. Gen Comp Endocrinol 201: 8-15.
http://dx.doi.Org/10.1016/i.ygcen.2014.03.019.
Woodward. DF; Farag. AM: Little. EE: Steadman. B; Yancik. R. (1991). Sensitivity of
greenback cutthroat trout to acidic pH and elevated aluminum. Trans Am Fish Soc 120:
34-42. http://dx.doi.org/10.1577/1548-8659(1991N)120<0034:SOGCTT>2.3.CO:2.
Wooldridge. PJ: Perring. AE; Bertram. TH: Flocke. FM; Roberts. JM; Singh. HB: Huev. LG:
Thornton. JA; Wolfe. GM; Murphy. JG: Fry. JL; Rollins. AW: LaFranchi. BW: Cohen.
RC. (2010). Total Peroxy Nitrates (Sigma PNs) in the atmosphere: The Thermal
Dissociation-Laser Induced Fluorescence (TD-LIF) technique and comparisons to
speciated PAN measurements. Atmos Meas Tech 3: 593-607.
http://dx.doi.org/10.5194/amt-3-593-2010.
Wooldridge. SA. (2009). Water quality and coral bleaching thresholds: formalising the linkage
for the inshore reefs of the Great Barrier Reef, Australia. Mar Pollut Bull 58: 745-751.
http://dx.doi.Org/10.1016/i.marpolbul.2008.12.013.
Worm. B; Barbier. EB; Beaumont. N: Duffy. JE; Folke. C: Halpern. BS: Jackson. JBC: Lotze.
HK: Micheli. F; Palumbi. SR; Sala. E; Selkoe. KA; Stachowicz. JJ; Watson. R. (2006).
Impacts of biodiversity loss on ocean ecosystem services. Science 314: 787-790.
http://dx.doi.Org/10.l 126/science.l 132294.
Wortman. E; Tomaszewski. T; Waldner. P; Schleppi. P; Thimonier. A: Eugster. W: Buchmann.
N: Sievering. H. (2012). Atmospheric nitrogen deposition and canopy retention
influences on photosynthetic performance at two high nitrogen deposition Swiss forests.
Tellus B Chem Phys Meteorol 64. http ://dx. doi. org/10.3402/tellusb.v64i0.17216.
WRI (World Resources Institute). (2000). World resources 2000-2001: People and ecosystems:
The fraying web of life. Washington, DC: World Recources Institute.
Wright. U; Reich. PB; Westobv. M; Ackerly. DP: Baruch. Z; Bongers. F; Cavender-Bares. J:
Chapin. T; Cornelissen. JHC; Diemer. M; Flexas. J: Gamier. E; Groom. PK; Gulias. J:
Hikosaka. K; Lamont. BB; Lee. T; Lee. W: Lusk. C: Midglev. JJ: Navas. ML: Niinemets.
U: Oleksyn. J: Osada. N: Poorter. H; Poot. P; Prior. L; Pyankov. VI; Roumet. C: Thomas.
SC; Tioelker. MG; Veneklaas. EJ; Villar. R. (2004). The worldwide leaf economics
spectrum. Nature 428: 821-827. http://dx.doi.org/10.1038/nature02403.
Wright. RF; Schindler. DW. (1995). Interaction of acid rain and global changes: Effects on
terrestrial and aquatic ecosystems. Water Air Soil Pollut 85: 89-99.
http://dx.doi.org/10.1007/BF0Q483691.
Wright. SHA; Berch. SM; Berbee. ML. (2009). The effect of fertilization on the below-ground
diversity and community composition of ectomycorrhizal fungi associated with western
hemlock (Tsuga heterophylla). Mycorrhiza 19: 267-276.
http://dx.doi.org/10.1007/s00572-008-Q218-x.
June 2018
R-243
DRAFT: Do Not Cite or Quote

-------
Wu, 0; Han. X; Wang. D, i; Zhao. F; Wang. D. (2017). Coal combustion related fine particulate
matter (PM2.5) induces toxicity in Caenorhabditis elegans by dysregulating microRNA
expression. Toxicology Research 6: 432-441. http://dx.doi.org/10.1039/c7txQ0107i.
Wu. W; Driscoll. CT. (2009). Application of the PnET-BGC - an integrated biogeochemical
model - to assess the surface water ANC recovery in the Adirondack region of New York
under three multi-pollutant proposals. J Hydrol 378: 299-312.
http://dx.doi.Org/10.1016/i.ihydrol.2009.09.035.
Wu. W: Driscoll. CT. (2010). Impact of climate change on three-dimensional dynamic critical
load functions. Environ Sci Technol 44: 720-726. http://dx.doi.org/10.1021/es90Q890t.
Wu. Z; Koch. GW: Diikstra. P; Bowker. MA: Hungate. BA. (2011). Responses of ecosystem
carbon cycling to climate change treatments along an elevation gradient. Ecosystems 14:
1066-1080. http://dx.doi.org/10.1007/slQ021-011-9464-4.
Wymore, AS: Coble, AA; Rodriguez-Cardona, B; Mcdowell, WH. (2016). Nitrate uptake across
biomes and the influence of elemental stoichiometry: A new look at LINXII. Global
Biogeochem Cycles 30: 1183-1191. http://dx.doi.org/10.1002/2016GB0Q5468.
Xia, J: Wan, S. (2008). Global response patterns of terrestrial plant species to nitrogen addition.
New Phytol 179: 428-439. http://dx.doi.org/10.1111/i. 1469-8137.2008.02488.x.
Xia, M; Talhelm, AF; Pregitzer, KS. (2015). Fine roots are the dominant source of recalcitrant
plant litter in sugar maple-dominated northern hardwood forests. New Phytol 208: 715-
726. http://dx.doi.org/10. Ill 1/nph. 13494.
Xiang, S, -R; Doyle, A: Holden, PA: Schimel, JP. (2008). Drying and rewetting effects on C and
N mineralization and microbial activity in surface and subsurface California grassland
soils. Soil Biol Biochem 40: 2281-2289. http://dx.doi.Org/10.1016/i.soilbio.2008.05.004.
Xiao, L; Xie, B; Liu, J: Zhang, H; Han, G: Wang, O; Liu, F. (2017). Stimulation of long-term
ammonium nitrogen deposition on methanogenesis by Methanocellaceae in a coastal
wetland. Sci Total Environ 595: 337-343.
http://dx.doi.Org/10.1016/i.scitotenv.2017.03.279.
Xing, J: Mathur, R; Pleim, J: Hogrefe, C: Gan, CM: Wong, DC: Wei, C: Gilliam, R; Pouliot, G.
(2015). Observations and modeling of air quality trends over 1990-2010 across the
Northern Hemisphere: China, the United States and Europe. Atmos Chem Phys 15: 2723-
2747. http://dx.doi.org/10.5194/acp-15-2723-2015.
Xing, J: Pleim, J: Mathur, R; Pouliot, G: Hogrefe, C: Gan, CM: Wei, C. (2013). Historical
gaseous and primary aerosol emissions in the United States from 1990 to 2010. Atmos
Chem Phys 13: 7531-7549. http://dx.doi.org/10.5194/acp-13-7531-2013.
Xing, Y; Bubier, J: Moore, T, im; Murphy, M; Basiliko, N; Wendel, S; Blodau, C. (2011). The
fate of N-15-nitrate in a northern peatland impacted by long term experimental nitrogen,
phosphorus and potassium fertilization. Biogeochemistry 103: 281-296.
http://dx.doi.org/10.1007/slQ533-010-9463-0.
June 2018
R-244
DRAFT: Do Not Cite or Quote

-------
Xiong, 0; Pan. K; Zhang. L, in; Wang. Y; Li. W. ei; He. X; Luo, H. (2016). Warming and
nitrogen deposition are interactive in shaping surface soil microbial communities near the
alpine timberline zone on the eastern Qinghai-Tibet Plateau, southwestern China. Appl
Soil Ecol 101: 72-83. http://dx.doi.Org/10.1016/i.apsoil.2016.01.011.
Xiong. S: Nilsson. C. (1999). The effects of plant litter on vegetation: A meta-analysis. J Ecol
87: 984-994. http://dx.doi.Org/10.1046/i.1365-2745.1999.00414.x.
Xiong. TT; Leveque. T; Austruv. A: Goix. S: Schreck. E; Dappe. V: Sobanska. S: Foucault. Y;
Dumat. C. (2014). Foliar uptake and metal(loid) bioaccessibility in vegetables exposed to
particulate matter. Environ Geochem Health 36: 897-909.
http://dx.doi.org/10.1007/slQ653-014-9607-6.
Xu. H; Yong. J: Xu. G. (2016a). Bioassessment of water quality status using a potential
bioindicator based on functional groups of planktonic ciliates in marine ecosystems. Mar
PollutBull 110: 409-414. http://dx.doi.Org/10.1016/i.marpolbul.2016.06.033.
Xu. L; Penner. JE. (2012). Global simulations of nitrate and ammonium aerosols and their
radiative effects. Atmos Chem Phys 12: 9479-9504. http://dx.doi.org/10.5194/acp-12-
9479-2012.
Xu, X; Meng, B; Zhang, C; Feng, X; Gu, C; Guo, J; Bishop, K; Xu, Z; Zhang, S; Qiu, G.
(2017a). The local impact of a coal-fired power plant on inorganic mercury and methyl-
mercury distribution in rice (Oryza sativa L.). Environ Pollut 223: 11-18.
http://dx.doi.Org/10.1016/i.envpol.2016.l 1.042.
Xu, X; Yang, F; Zhao, L; Yan, X. (2016b). Seawater acidification affects the physiological
energetics and spawning capacity of the Manila clam Ruditapes philippinarum during
gonadal maturation. Comp Biochem Physiol AMol Integr Physiol 196: 20-29.
http://dx.doi.Org/10.1016/i.cbpa.2016.02.014.
Xu, Z; Ren, H; Li, M, aiHe; Brunner, I; Yin, J; Liu, H; Kong, D; Lu, XT, ao; Sun, T, ao; Cai, J;
Wang, R; Zhang, Y; He, P; Han, X: Wan, S: Jiang, Y. (2017b). Experimentally increased
water and nitrogen affect root production and vertical allocation of an old-field grassland.
Plant Soil 412: 369-380. http://dx.doi.org/10.1007/slll04-016-3071-2.
Xun, L; Campbell, NER; Rudd, JWM. (1987). Measurements of specific rates of net methyl
mercury production in the water column and surface sediments of acidified and
circumneutral lakes. Can J Fish Aquat Sci 44: 750-757. http://dx.doi.org/10.1139/f87-
091.
Yamulki, S; Harrison, RM; Goulding, KWT; Webster, CP. (1997). N20, NO and N02 fluxes
from a grassland: Effect of soil pH. Soil Biol Biochem 29: 1199-1208.
http://dx.doi. org/10.1016/S003 8-0717(97^)00032-1.
Yan, ND; Keller, W: Somers, KM: Pawson, TW: Girard, RE. (1996a). Recovery of crustacean
zooplankton communities from acid and metal contamination: Comparing manipulated
and reference lakes. Can J Fish Aquat Sci 53: 1301-1327. http://dx.doi.org/10.1139/f96-
065.
June 2018
R-245
DRAFT: Do Not Cite or Quote

-------
Yan, ND; Welsh. PG; Lin. H; Taylor. DJ: Filion, J. -M. (1996b). Demographic and genetic
evidence of the long-term recovery of Daphnia galeata mendotae (Crustacea: Daphniidae)
in Sudbury lakes following additions of base: The role of metal toxicity. Can J Fish
Aquat Sci 53: 1328-1344. http://dx.doi.Org/10.l 139/cifas-53-6-1328.
Yan. X: Shi. W. enZ; Zhao. W. enJ; Luo. N. aN. (2014). Impact of aerosols and atmospheric
particles on plant leaf proteins. Atmos Environ 88: 115-122.
http://dx.doi.ore/10.1016/i.atmosenv.2014.01.044.
Yanai. RD; Vadeboncoeur. MA: Hamburg. SP; Arthur. MA: Fuss. CB; Groffinan. PM; Siccama.
TG: Driscoll. CT. (2013). From missing source to missing sink: Long-term changes in the
nitrogen budget of a northern hardwood forest. Environ Sci Technol 47: 11440-11448.
http://dx.doi.org/10.1021/es4025723.
Yang. CH. (1998). Electrochemical polymerization of aniline and toluidines on a thermally
prepared Pt electrode. Journal of Electroanalytical Chemistry 459: 71-89.
http://dx.d0i.0rg/l 0.1016/S0022-0728(98X)0292-7.
Yang. R; Rui. Q. i; Kong. L; Zhang. N. an: Li. Y. u: Wang. X: Tao. J: Tian. P; Ma. Y. an: Wei. J:
Li. G: Wang. D. (2016). Metallothioneins act downstream of insulin signaling to regulate
toxicity of outdoor fine particulate matter (PM2.5) during Spring Festival in Beijing in
nematode Caenorhabditis elegans. Toxicology Research 5: 1097-1105.
http://dx.doi.org/10.1039/c6txQ0022c.
Yang. R; Zhao. Y; Yu. X: Lin. Z; Xi. Z; Rui. O, i; Wang. D. (2015). Insulin signaling regulates
the toxicity of traffic-related PM2.5 on intestinal development and function in nematode
Caenorhabditis elegans. Toxicology Research 4: 333-343.
http://dx.doi.org/10.1039/c4tx00131a.
Yang. WH: Ryals. RA; Cusack. DF; Silver. WL. (2017). Cross-biome assessment of gross soil
nitrogen cycling in California ecosystems. Soil Biol Biochem 107: 144-155.
http://dx.doi.Org/10.1016/i.soilbio.2017.01.004.
Yang. X: Richardson. TK; Jain. AK. (2010). Contributions of secondary forest and nitrogen
dynamics to terrestrial carbon uptake. Biogeosciences 7: 3041-3050.
http://dx.doi.org/10.5194/bg-7-3041-2010.
Yang. XE; Jin. XF; Feng. Y; Islam. E. (2005). Molecular mechanisms and genetic basis of heavy
metal tolerance/hyperaccumulation in plants. J Integr Plant Biol 47: 1025-1035.
http://dx.doi.org/10.1111/i. 1744-7909.2005.00144.x.
Yang. Y: He. Z; Lin. Y; Phlips. EJ; Yang. J: Chen. G: Stoffella. PJ: Powell. CA. (2008).
Temporal and spatial variations of nutrients in the Ten Mile Creek of south Florida, USA
and effects on phytoplankton biomass. J Environ Monit 10: 508-516.
http://dx.doi.org/10.1039/b715026a.
Yatavelli, RLN; Fahrni, JK; Kim, M; Crist, KC; Vickers, CD: Winter, SE; Connell, DP. (2006).
Mercury, PM2.5 and gaseous co-pollutants in the Ohio River Valley region: Preliminary
results from the Athens supersite. Atmos Environ 40: 6650-6665.
http://dx.doi.Org/10.1016/i.atmosenv.2006.05.072.
June 2018
R-246
DRAFT: Do Not Cite or Quote

-------
Ye. C; Zhou. X; Pu, D; Stutz, J; Festa, J; Spolaor, M; Tsai, C; Cantrell, C; Mauldin, RL;
Campos. T; Weinheimer. A; Hornbrook. RS; Apel. EC; Guenther. A; Kaser. L; Yuan. B;
Karl, T; Haggerty, J; Hall, S; Ullmann, K; Smith, JN; Ortega, J; Knote, C. (2016). Rapid
cycling of reactive nitrogen in the marine boundary layer. Nature 532: 489-491.
http://dx.doi.org/10.1038/naturel7195.
Ye, R; Wright, AL; Orem, WH; McCrav, JM. (2010). Sulfur distribution and transformations in
Everglades agricultural area soil as influenced by sulfur amendment. Soil Sci 175: 263-
269. http://dx.doi.org/10.1097/SS.0b013e3181el6168.
Yin, T, -F; Zheng, L, -L, C.,ao, G.,-M; Song, M, -H; Yu, F, -H. (2017). Species-specific
phenological responses to long-term nitrogen fertilization in an alpine meadow. Journal
of Plant Ecology 10: 301-309. http://dx.doi.org/10.1093/ipe/rtw026.
Ying, J: Li, X: Wang, N: Lan, Z; He, J: Bai, Y. (2017). Contrasting effects of nitrogen forms and
soil pH on ammonia oxidizing microorganisms and their responses to long-term nitrogen
fertilization in atypical steppe ecosystem. Soil Biol Biochem 107: 10-18.
http://dx.doi.Org/10.1016/i.soilbio.2016.12.023.
Yoshida, LC: Allen, EB. (2001). Response to ammonium and nitrate by a mycorrhizal annual
invasive grass and native shrub in southern California. Am J Bot 88: 1430-1436.
http://dx.doi.org/10.2307/355845Q.
Young, CS: Gobler, CJ. (2016). Ocean Acidification Accelerates the Growth of Two Bloom-
Forming Macroalgae. PLoS ONE 11: e0155152.
http://dx.doi.org/10.1371/iournal.pone.0155152.
Yu, F; Zou, J: Hua, Y; Zhang, S: Liu, G: Zhu, D. (2015). Transformation of external sulphate
and its effect on phosphorus mobilization in Lake Moshui, Wuhan, China. Chemosphere
138: 398-404. http://dx.doi.Org/10.1016/i.chemosphere.2015.06.067.
Yu, RQ; Adatto, I; Montesdeoca, MR; Driscoll, CT; Hines, ME; Barkay, T. (2010). Mercury
methylation in Sphagnum moss mats and its association with sulfate-reducing bacteria in
an acidic Adirondack forest lake wetland. FEMS Microbiol Ecol 74: 655-668.
http://dx.doi.Org/10.llll/i.1574-6941.2010.00978.x.
Yu, RQ; Flanders, J. R.; Mack, EE; Turner, R; Mirza, MB; Barkay, T. (2012). Contribution of
coexisting sulfate and iron reducing bacteria to methylmercury production in freshwater
river sediments. Environ Sci Technol 46: 2684-2691.
http://dx.doi.org/10.1021/es2033718.
Yu, X, ue; Driscoll, CT; Montesdeoca, M; Evers, D; Duron, M; Williams, K; Schoch, N;
Kamman, NC. (2011). Spatial patterns of mercury in biota of Adirondack, New York
lakes. Ecotoxicology 20: 1543-1554. http://dx.doi.org/10.1007/slQ646-011-0717-v.
Yuan, LL; Pollard, AI; Pather, S; Oliver, JL; D'Anglada, L. (2014). Managing microcystin:
Identifying national-scale thresholds for total nitrogen and chlorophyll a. Freshw Biol 59:
1970-1981. http://dx.doi.org/10.1111/fwb. 12400.
Yuan, X, ia; Knelman, JE; Gasarch, E, ve; Wang, D; Nemergut, PR; Seastedt, TR. (2016). Plant
community and soil chemistry responses to long-term nitrogen inputs drive changes in
alpine bacterial communities. Ecology 97: 1543-1554. http://dx.doi.org/10.1890/15-
1160.1.
June 2018
R-247
DRAFT: Do Not Cite or Quote

-------
Yue, K; Fornara, DA; Yang. W; Peng. Y; Li. Z; Wu, F; Peng. C. (2017). Effects of three global
change drivers on terrestrial C:N:P stoichiometry: a global synthesis. Global Change Biol
23: 2450-2463. http://dx.doi.Org/10.l 111/gcb. 13569.
Yue. K; Peng. Y; Peng. C: Yang. W: Peng. X: Wu. F. (2016). Stimulation of terrestrial
ecosystem carbon storage by nitrogen addition: A meta-analysis. Sci Rep 6: 19895.
http://dx.doi.org/10.1038/srepl9895.
Zaehle. S. (2013). Terrestrial nitrogen-carbon cycle interactions at the global scale. Philos Trans
R Soc Lond B Biol Sci 368: 20130125. http://dx.doi.org/10.1098/rstb.2013.Q125.
Zaehle, S; Dalmonech, D. (2011). Carbon-nitrogen interactions on land at global scales: Current
understanding in modelling climate biosphere feedbacks. Current Opinion in
Environmental Sustainability 3: 311-320. http://dx.doi.Org/10.1016/i.cosust.2011.08.008.
Zaehle. S: Friedlingstein. P; Friend. AD. (2010). Terrestrial nitrogen feedbacks may accelerate
future climate change. Geophys Res Lett 37. http://dx.doi.org/10.1029/2009GLQ41345.
Zaiac. K; Blodau. C. (2016). The fate of N-15-nitrate in mesocosms from five European
peatlands differing in long-term nitrogen deposition rate. Biogeosciences 13: 707-722.
http://dx.doi.org/10.5194/bg-13-707-2016.
Zak. PR: Freedman. ZB; Upchurch. RA; Steffens. M; Kogel-Knabner. I. (2017). Anthropogenic
N deposition increases soil organic matteraccumulation without altering its biochemical
composition. Global Change Biol 23: 933-944. http://dx.doi.Org/10.l 111/gcb. 13480.
Zak. PR: Holmes. WE: Burton. AJ; Pregitzer. KS: Talhelm. AF. (2008). Simulated atmospheric
N03 deposition increases soil organic matter by slowing decomposition. Ecol Appl 18:
2016-2027. http://dx.doi.Org/10.1890/07-1743.l.
Zak, PR; Pregitzer, KS: Burton, AJ: Edwards, IP: Kellner, H. (2011). Microbial responses to a
changing environment: implications for the future functioning of terrestrial ecosystems.
Fungal Ecology 4: 386-395. http://dx.doi.Org/10.1016/i.funeco.2011.04.001.
Zaldivar, JM; Cardoso, AC: P, V: Newton, A: de Wit, R; Ibanez, C: Reizopoulou, S: Francesca,
S; Razinkovas, A; Basset, A; Holmer, M; Murray, N. (2008). Eutrophication in
transitional waters: An overview. Transitional Water Monographs 1: 1-78.
http://dx.doi.org/10.1285/il8252273v2nlpl.
Zamin, TJ; Bret-Harte, MS: Grogan, P. (2014). Evergreen shrubs dominate responses to
experimental summer warming and fertilization in Canadian mesic low arctic tundra. J
Ecol 102: 749-766. http://dx.doi.Org/10.l 111/1365-2745.12237.
Zamin, TJ: Grogan, P. (2012). Birch shrub growth in the low Arctic: The relative importance of
experimental warming, enhanced nutrient availability, snow depth and caribou exclusion.
Environ Res Lett 7. http://dx.doi.Org/10.1088/1748-9326/7/3/034027.
Zamora, LM; Prospero, JM; Hansell, PA. (2011). Organic nitrogen in aerosols and precipitation
at Barbados and Miami: Implications regarding sources, transport and deposition to the
western subtropical North Atlantic. J Geophys Res Atmos 116.
http://dx.doi.org/10.1029/2011JP01566Q.
June 2018
R-248
DRAFT: Do Not Cite or Quote

-------
Zanchi. G; Belvazid. S; Akselsson. C; Yu. L; Bishop. K; Kohler. SJ; Grip. H. (2016). A
hydrological concept including lateral water flow compatible with the biogeochemical
model ForSAFE. Hydrology 3:11. http://dx.doi.org/10.3390/hydrology3010011.
Zarnoch. CB; Hoellein. TJ; Furman. BT; Peterson. BJ. (2017). Eel grass meadows, Zostera
marina (L.), facilitate the ecosystem service of nitrogen removal during simulated
nutrient pulses in Shinnecock Bay, New York, USA. Mar Pollut Bull 124: 376-387.
http://dx.doi.ore/10.1016/j.marpolbul.2017.07.061.
Zavaleta. ES: Pasari. J. aeR; Hulvev. KB: Tilman. GD. (2010). Sustaining multiple ecosystem
functions in grassland communities requires higher biodiversity. Proc Natl Acad Sci USA
107: 1443-1446. http://dx.doi.org/10.1073/pnas.09068291Q7.
Zavaleta. ES: Shaw. MR: Chiariello. NR; Moonev. HA: Field. CB. (2003). Additive effects of
simulated climate changes, elevated C02, and nitrogen deposition on grassland diversity.
Proc Natl Acad Sci USA 100: 7650-7654. http://dx.doi.org/10.1073/pnas.093273410Q.
Zechmeister-Boltenstern. S: Michel. K; Pfeffer. M. (2011). Soil microbial community structure
in European forests in relation to forest type and atmospheric nitrogen deposition. Plant
Soil 343: 37-50. http://dx.doi.org/10.1007/slllQ4-010-0528-6.
Zehnder, CB; Hunter, MP. (2008). Effects of nitrogen deposition on the interaction between an
aphid and its host plant. Ecol Entomol 33: 24-30. http://dx.doi.Org/10.l 11 l/j.1365-
2311.2007.00945.x.
Zeng. J. un; Liu. X: Song. L; Lin. X: Zhang. H; Shen. C: Chu. H. (2016). Nitrogen fertilization
directly affects soil bacterial diversity and indirectly affects bacterial community
composition. Soil Biol Biochem 92: 41-49.
http://dx.doi.Org/10.1016/i.soilbio.2015.09.018.
Zerbe. P; Chiang. A: Yuen. M; Hamberger. B; Hamberger. B; Draper. JA; Britton. R; Bohlmann.
J. (2012). Bifunctional cis-abienol synthase from Abies balsamea discovered by
transcriptome sequencing and its implications for diterpenoid fragrance production. J
Biol Chem 287: 12121-12131. http://dx.doi.org/10.1074/ibc.Mlll.317669.
Zetterberg. T; Kohler. SJ: Lofgren. S. (2014). Sensitivity analyses of MAGIC modelled
predictions of future impacts of whole-tree harvest on soil calcium supply and stream
acid neutralizing capacity. Sci Total Environ 494-495: 187-201.
http://dx.doi.Org/10.1016/i.scitotenv.2014.06.l 14.
Zhai. J: Driscoll. CT; Sullivan. TJ: Cosby. BJ. (2008). Regional application of the PnET-BGC
model to assess historical acidification of Adirondack Lakes. Water Resour Res 44.
http://dx.doi.org/10.1029/2006WR0Q5532.
Zhalnina. K; Dias. R; de Quadros. PD; Davis-Richardson. A: Camargo. FAQ: Clark. IM;
McGrath. SP; Hirsch. PR: Triplett. EW. (2015). Soil pH determines microbial diversity
and composition in the park grass experiment. Microb Ecol 69: 395-406.
http://dx.doi.org/10.1007/s00248-014-053Q-2.
Zhang. D: Hui. D: Luo. Y; Zhou. G. (2008). Rates of litter decomposition in terrestrial
ecosystems: Global patterns and controlling factors. Journal of PI ant Ecology 1: 85-93.
http://dx.doi.org/10.1093/ipe/rtn002.
June 2018
R-249
DRAFT: Do Not Cite or Quote

-------
Zhang. J. an; Hudson. J; Neal. R; Sereda. J; Clair. T; Turner. M; Jeffries. D; Dillon. P; Molot L;
Somers. K; Hesslein. R. ay. (2010). Long-term patterns of dissolved organic carbon in
lakes across eastern Canada: Evidence of a pronounced climate effect. Limnol Oceanogr
55: 30-42. http://dx.doi.Org/10.4319/lo.2010.55.l.0030.
Zhang. J: Li. H; Shen. H; Chen. Y; Fang. J: Tang. Z. (2015a). Effects of nitrogen addition on
nitrogen resorption in temperate shrublands in northern China. PLoS ONE 10: e0130434.
http://dx.doi.org/10.1371/iournal.pone.013Q434.
Zhang. L; Brook. JR; Vet. R. (2003). Evaluation of a non-stomatal resistance parameterization
for S02 dry deposition. Atmos Environ 37: 2941-2947. http://dx.doi.org/10.lQ16/S1352-
2310(03 W0268-1.
Zhang. L; Gong. S: Padro. J: Barrie. L. (2001). A size-segregated particle dry deposition scheme
for an atmospheric aerosol module. Atmos Environ 35: 549-560.
http ://dx. doi .org/10.1016/S13 52-2310(00^)00326-5.
Zhang. L; Jacob. DJ: Downey. NV: Wood. DA: Blewitt. D; Carouge. CC: van Donkelaar. A:
Jones. DBA: Murray. LT; Wang. Y. (201 la). Improved estimate of the policy-relevant
background ozone in the United States using the GEOS-Chem global model with 1/2 2/3
horizontal resolution over North America. Atmos Environ 45: 6769-6776.
http://dx.doi.Org/10.1016/i.atmosenv.2011.07.054.
Zhang. L; Jacob. DJ: Knipping. EM: Kumar. N: Munger. JW: Carouge. CC: van Donkelaar. A:
Wang. YX; Chen. D. (2012a). Nitrogen deposition to the United States: Distribution,
sources, and processes. Atmos Chem Phys 12: 4539-4554. http://dx.doi.org/ 10.5194/acp-
12-4539-2012.
Zhang. Q: Wang. YP; Pitman. AJ; Dai. YJ. (201 lb). Limitations of nitrogen and phosphorous on
the terrestrial carbon uptake in the 20th century. Geophys Res Lett 38.
http://dx.doi.org/10.1029/2011GL049244.
Zhang. O: Xie. J: Lvu. M: Xiong. D: Wang. J: Chen. Y: Li. Y: Wang. M: Yang. Y. (2017).
Short-term effects of soil warming and nitrogen addition on the N:P stoichiometry of
Cunninghamia lanceolata in subtropical regions. Plant Soil 411: 395-407.
http://dx.doi.org/10.1007/slllQ4-016-3037-4.
Zhang. T; Guo. R; Gao. S: Guo. J: Sun. W. (2015b). Responses of plant community composition
and biomass production to warming and nitrogen deposition in a temperate meadow
ecosystem. PLoS ONE 10: e0123160. http://dx.doi.org/10.1371/iournal.pone.0123160.
Zhang. W: Shen. W: Zhu. S: Wan. S: Luo. Y; Yan. J: Wang. K; Liu. L; Dai. H; Li. P; Dai. K;
Zhang. W: Liu. Z; Wang. F; Kuang. Y; Li. Z; Lin. Y; Rao. X: Li. J: Zou. B; Cai. X: Mo.
J: Zhao. P; Ye. O: Huang. J: Fu. S. (2015c). CAN canopy addition of nitrogen better
illustrate the effect of atmospheric nitrogen deposition on forest ecosystem? Sci Rep 5:
11245. http://dx.doi.org/10.1038/srepl 1245.
Zhang. X: Tan. Y; Li. A: Ren. T; Chen. S: Wang. L; Huang. J. (2015d). Water and nitrogen
availability co-control ecosystem C02 exchange in a semiarid temperate steppe. Sci Rep
5. http://dx.doi.org/10.1038/srepl5549.
June 2018
R-250
DRAFT: Do Not Cite or Quote

-------
Zhang. XH; Zhu, YG; Chen. BP; Lin. AJ; Smith. SE; Smith. FA. (2005). Arbuscular
mycorrhizal fungi contribute to resistance of upland rice to combined metal
contamination of soil. J Plant Nutr 28: 2065-2077.
http://dx.doi.org/10.1080/0190416050032Q871.
Zhang. Y; Feng. J: Isbell. F; Lti. X: Han. X. (2015e). Productivity depends more on the rate than
the frequency of N addition in a temperate grassland. Sci Rep 5: 12558.
http://dx.doi.org/10.1038/srepl2558.
Zhang. Y; Loreau. M; Lu. X: He. N: Zhang. G: Han. X. (2016a). Nitrogen enrichment weakens
ecosystem stability through decreased species asynchrony and population stability in a
temperate grassland. Global Change Biol 22: 1445-1455.
http://dx.doi.org/10. Ill 1/gcb. 13140.
Zhang. Y; Lii. X: Isbell. F; Stevens. C: Han. X: He. N: Zhang. G: Yu. Q: Huang. J: Han. X.
(2014). Rapid plant species loss at high rates and at low frequency of N addition in
temperate steppe. Global Change Biol 20: 3520-3529.
http://dx.doi.org/10. Ill 1/gcb. 12611.
Zhang. Y; Stevens. CJ: Lu. X: He. N: Huang. J: Han. X. (2016b). Fewer new species colonize at
low frequency N addition in a temperate grassland. Funct Ecol 30: 1247-1256.
http://dx.doi.org/10.1111/1365-2435.12585.
Zhang. Y. u: Chen. H. anYH; Reich. PB. (2012b). Forest productivity increases with evenness,
species richness and trait variation: a global meta-analysis. J Ecol 100: 742-749.
http://dx.doi.org/10.1111/i. 1365-2745.2011.01944.x.
Zhang. Y; Xu. X: Li. Y; Huang. L; Xie. X: Dong. J: Yang. S. (2016c). Effects of Spartina
alterniflora invasion and exogenous nitrogen on soil nitrogen mineralization in the coastal
salt marshes. Ecol Eng 87: 281-287. http://dx.doi.Org/10.1016/i.ecoleng.2015.12.003.
Zhao, C; Zhu, L; Liang, J, in; Yin, H; Yin, C; Li, D; Zhang, N; Liu, Q. (2014a). Effects of
experimental warming and nitrogen fertilization on soil microbial communities and
processes of two subalpine coniferous species in eastern Tibetan Plateau, China. Plant
Soil 382: 189-201. http://dx.doi.org/10.1007/sl 1104-014-2153-2.
Zhao, Y; Lin, Z; Jia, R; Li, G; Xi, Z; Wang, D. (2014b). Transgenerational effects of traffic-
related fine particulate matter (PM2.5) on nematode Caenorhabditis elegans. J Hazard
Mater 274: 106-114. http://dx.doi.Org/10.1016/i.ihazmat.2014.03.064.
Zheng, S: Gu, B; Zhou, Q: Li, Y. (2013). Variations of mercury in the inflow and outflow of a
constructed treatment wetland in south Florida, USA. Ecol Eng 61: 419-425.
http://dx.doi.Org/10.1016/i.ecoleng.2013.10.015.
Zheng, Y; Hou, L; Newell, S: Liu, M, in: Zhou, J: Zhao, H, ui; You, L; Cheng, X. (2014).
Community dynamics and activity of ammonia-oxidizing prokaryotes in intertidal
sediments of the Yangtze Estuary. Appl Environ Microbiol 80: 408-419.
http://dx.doi.org/10.1128/AEM.03035-13.
Zhou, L; Zhou, X: Zhang, B; Lu, M; Luo, Y; Liu, L; Li, B. (2014a). Different responses of soil
respiration and its components to nitrogen addition among biomes: A meta-analysis.
Global Change Biol 20: 2332-2343. http://dx.doi.org/10. Ill 1/gcb. 12490.
June 2018
R-251
DRAFT: Do Not Cite or Quote

-------
Zhou. 0; Driscoll, CT; Moore. SE; Kulp, MA; Renfro, JR; Schwartz. JS; Cat M; Lynch. JA.
(2015a). Developing critical loads of nitrate and sulfate deposition to watersheds of the
Great Smoky Mountains National Park, USA. Water Air Soil Pollut 226.
http://dx.doi.org/10.1007/sll270-015-25Q2-7.
Zhou. O; Driscoll. CT; Sullivan. TJ. (2015b). Responses of 20 lake-watersheds in the
Adirondack region of New York to historical and potential future acidic deposition. Sci
Total Environ 511: 186-194. http://dx.doi.Org/10.1016/i.scitotenv.2014.12.044.
Zhou. Q; Driscoll. CT; Sullivan. TJ; Pourmokhtarian. A. (2015c). Factors influencing critical and
target loads for the acidification of lake-watersheds in the Adirondack region of New
York. Biogeochemistry 124: 353-369. http://dx.doi.org/10.1007/slQ533-015-0102-7.
Zhou. XB; Zhang. YM; Yin. BF. (2016). Divergence in physiological responses between
cyanobacterial and lichen crusts to a gradient of simulated nitrogen deposition. Plant Soil
399: 121-134. http://dx.doi.org/10.1007/slll04-015-2687-v.
Zhou. Y; Scavia. D; Michalak. AM. (2014b). Nutrient loading and meteorological conditions
explain interannual variability of hypoxia in Chesapeake Bay. Limnol Oceanogr 59: 373-
384. http://dx.doi.Org/10.4319/lo.2014.59.2.0373.
Zhu. J; Wang. Q; He. N; Smith. MP; Elser. JJ; Du. J; Yuan. G; Yu. G; Yu. Q. (2016a).
Imbalanced atmospheric nitrogen and phosphorus depositions in China: Implications for
nutrient limitation. Jour Geo Res: Biog 121: 1605-1616.
http://dx.doi. org/10.1002/2016JG003 393.
Zhu, W; Wang, J; Zhang, Z; Ren, F; Chen, L; He, JS. (2016b). Changes in litter quality induced
by nutrient addition alter litter decomposition in an alpine meadow on the Qinghai-Tibet
Plateau. Sci Rep 6: 34290. http://dx.doi.org/10.1038/srep34290.
Ziegler, I. (1973). The effect of air-polluting gases on plant metabolism. In F Coulston; F Korte
(Eds.), Environmental Quality and Safety: Global Aspects of Chemistry, Toxicology and
Technology as Applied to the Environment (pp. 182-208). New York, NY: Thieme,
Academic Press.
Zimmer, D; Baum, C; Leinweber, P; Hrynkiewicz, K; Meissner, R. (2009). Associated bacteria
increase the phytoextraction of cadmium and zinc from a metal-contaminated soil by
mycorrhizal willows. Int J Phytoremediation 11: 200-213.
http://dx.doi.org/10.1080/152265108Q2378483.
Zong, N; Shi, P; Song, M; Zhang, X; Jiang, J; Chai, X, i. (2016). Nitrogen Critical Loads for an
Alpine Meadow Ecosystem on the Tibetan Plateau. Environ Manage 57: 531-542.
http://dx.doi.org/10.1007/sQ0267-015-0626-6.
June 2018
R-252
DRAFT: Do Not Cite or Quote

-------