IvvEPA
United States
Environmental Protection
Agency
Office of Environmental
Processes and
Effects Research
Washington DC 20460
EPA/600/R-92/186
Research and Development
September 1992
Direct/Delayed Response
Project: Future Effects of
Long-Term Sulfur Deposition
on Stream Chemistry in the
Mid-Appalachian Region of
the Eastern United States
-------
-------
EPA/600/R-92/186
September 1992
Direct/Delayed Response Project:
Future Effects of Long-Term Sulfur Deposition
on Stream Chemistry in the
Mid-Appalachian Region of the
Eastern United States
By
M.R.Church, P.W.Shaffer, K.W.Thornton, D.L.Cassell, C.I. Liff,
M.G. Johnson, DA Lammers, J.J. Lee, G.R. Holdren, J.S. Kern,
LH.Liegel, S.M. Pierson, D.L.Stevens, B.P. Rochelle, R.S.Turner
A Contribution to the
National Acid Precipitation Assessment Program
U.S. Environmental Protection Agency
Office of Research and Development, Washington DC 20460
Environ mental Research Laboratory, Corvallis, Oregon 97333
-------
NOTICE
The information in this document has been funded wholly (or in part) by the U.S. Environmental
Protection Agency. It has been subjected to the Agency's peer and administrative review, and it has
been approved for publication as an EPA document. Mention of trade names or commercial products
does not constitute endorsement or recommendation for use.
-------
CONTENTS
Section
Page
Notice "
Figures viii
Plates • xiii
Tables
Primary Contributors •
Acknowledgments x*
Preface xxiii
Preface to the Executive Summary xxiv
1 EXECUTIVE SUMMARY 1
1.1 INTRODUCTION 1
1.1.1 Project Background 1
1.1.2 Primary Objectives 2
1.1.3 Study Regions 2
1.1.4 Time Frames of Concern 3
1.2 PROCESSES OF ACIDIFICATION 3
1.2.1 Sulfur Retention ., 3
1.2.2 Base Cation Supply 5
1.3 GENERAL APPROACH 5
1.3.1 Watershed Selection 7
1.3.2 Soil Survey 7
1.3.3 Other Regional Datasets 7
1.3.4 Scenarios of Atmospheric Deposition 7
1.3.5 Data Analysis 7
1.4 RESULTS 9
1.4.1 Retention of Atmospherically Deposited Sulfur 10
1.4.1.1 Current Retention • "1°
1.4.1.2 Projected Retention 10
1.4.2. Base Cation Supply 13
1.4.2.1 Current Control 13
1.4.2.2 Future Effects ' 14
1.4.3 Integrated Effects of Surface Water ANC 14
1.5 SUMMARY DISCUSSION 20
1.6 REFERENCES 25
2 INTRODUCTION TO THE DIRECT/DELAYED RESPONSE PROJECT 31
2.1 PROJECT BACKGROUND 31
2.2 PRIMARY OBJECTIVES 32
2.3 STUDY REGIONS 32
2.4 TIME FRAMES OF CONCERN S4
2.5 PROJECT PARTICIPANTS 34
2.6 REPORTING 34
3 PROCESSES OF ACIDIFICATION " 37
3.1 INTRODUCTION 37
3.2 THE MOBILE ANION PARADIGM - A CONCEPTUAL MODEL OF ACIDIFICATION 37
3.3 ANION MOBILITY • • • 38
3.3.1 Sulfate 38
3.3.2 Other Anions 40
III
-------
9.2.8.1 Model Description 161
9.2.8.2 Data Sources 162
9.2.3.3 Model Assumptions and Limitations 163
9.2.8.4 Adsorption Data 165
9.2.8.5 Calibration and Evaluation of Aggregated Data 167
9.2.8.6 Target Populations for Model Projections 168
9.2.4 Results 169
9.2.4.1. Summary of Soil Adsorption Data 169
9.2.4.2. Evaluation of Aggregated Data 169
9.2.4.3. Base Year - Measured Sulfate and Model Projections 174
9.2.4.4. Model Results - Projections for Current Sulfur Deposition 180
9.2.4.5. Model Results - Projections for Scenarios of Reduced Future Deposition ... 186
9.2.5. Summary 189
9.3 EFFECT OF CATION EXCHANGE AND WEATHERING ON SYSTEM RESPONSE 193
9.3.1 Introduction 193
9.3.t.1 Level II Hypotheses 194
9.3.1.2 Approach 196
9.3.2 The BIoom-Grigal Model 198
9.3.2.1 Model Formulation 198
9.3.2.2 Assumptions 201
9.3.2.3 Limitations 204
9.3.2.4 Model Inputs 204
9.3.2.5 Model Outputs 206
9.3.2.6 Data Sources " 206
9.3.2.7 Results 208
9.3.2.8 Regional Comparisons 226
9.3.3 Reuss Model 232
9.3.S.1 Model Description 232
9.3.3.2 Model Formulation 235
9.3.3.3 Assumptions 235
9.3.3.4 Limitations 235
9.3.3.5 Model Inputs 236
9.3.3.6 Results - ANC Projections 236
9.3.3.7 Conclusion 245
9.4 SUMMARY 248
10 LEVEL III ANALYSES - DYNAMIC WATERSHED MODELLING 251
10.1 INTRODUCTION 251
10.2 MODEL. OF ACIDIFICATION OF GROUNDWATER IN CATCHMENTS (MAGIC) 253
10.3 OPERATIONAL ASSUMPTIONS 255
10.4 MODELLING DATASETS 255
10.4.1 Meteorological/Deposition Data 257
10.4.2 DDRP Runoff Estimation 257
10.4.2.1 Annual Runoff 257
10.4.2.2 Monthly Runoff 257
10.4.3 Watershed and Stream Morphometry Data 259
10.4.4 Soil/Stream Physical and Chemical Data 259
10.4.5 Other Data 259
10.5 CALIBRATION OF MAGIC '.'. 264
10.5.1 Calibration Procedure 264
10.5.2 Calibration/Confirmation Studies 264
10.5.2.1 Study Watersheds 264
10.5.2.2 Results 265
10.6 MODEL PROJECTIONS FOR MID-APPALACHIAN REGION 271
10.6.1 General Approach 271
VI
-------
10.6.2 Target Population for MAGIC Projections 274
10.6.3 Regional Calibration 276
10.6.4 Mid-Appalachian Projections 281
10.6.4.1 Projected Values of ANC, pH, Sulfate and Calcium Plus Magnesium 281
10.6.4.2 Projected Rates of Change 293
10.6.4.3 Comparison With ILWAS Projections for Two Watersheds 305
10.7 DISCUSSION 305
10.7.1 Projections of Future Changes in Acid-Base Surface Water Chemistry 305
10.7.2 Systems of Interest 306
10.7.3 Relative Magnitude and Direction of Change 308
10.7.4 Relationships Among Model Projections and Watershed Processes 309
10.7.4.1 Sulfur Retention Processes 309
10.7,4.2 Base Cation Supply 310
10.7.4.3 Nitrogen Dynamics 311
10.7.4.4 Land Use Changes 312
10.7.4.5 Organic Acidification 312
10.7.5 Other Considerations 312
10.7.5.1 Data Aggregation 312
10.7.5.2 Effects on Upstream Versus Downstream Nodes 313
10.8 REGIONAL COMPARISON: NE, M-APP, SBRP , 315.
10.9 CONCLUSIONS FROM LEVEL III ANALYSES 316
11 SUMMARY OF RESULTS 317
11.1 RETENTION OF ATMOSPHERICALLY DEPOSITED SULFUR 317
11.1.1 Current Retention 317
11.1.2 Projected Retention 317
11.2 BASE CATION SUPPLY 319
11.2.1 Current Control 319
11.2.2 Future Effects 321
11.3 INTEGRATED EFFECTS OF SURFACE WATER ANC 322
11.4 SUMMARY DISCUSSION 327
12 REFERENCES 333
13 GLOSSARY 355
13.1 ABBREVIATIONS AND SYMBOLS 355
13.1.1 Abbreviations .. 355
13.1.2 Symbols 357
13.2 DEFINITIONS 360
VII
-------
FIGURES
PAGE
1-1. Steps of the Direct/Delayed Response Project (DDRP) approach 6
1-2. Sulfur deposition scenarios for the M-APP Region 8
1-3. Percentage of stream reaches of the DDRP M-APP modelling target population
projected by the MAGIC model to have ANC < 0 peq/L at the lower reach nodes 21
1-4. Length of stream reaches of the DDRP M-APP modelling target population
projected by the MAGIC model to have ANC < 0 peq/Lat the lower reach nodes 22
1-5. Percentage of stream reaches of the DDRP M-APP modelling target population
projected by the MAGIC model to have ANC < 50 fieq/L at the lower reach nodes 23
1-6. Length of stream reaches of the DDRP M-APP modelling target population
projected by the MAGIC model to have ANC < 50 ^eq/L at the lower reach nodes 24
2-1. Activities of the Aquatic Effects Research Program within the National Acid Precipitation
Assessment Program 33
4-1. Steps in the Direct/Delayed Response Project (DDRP) approach 46
5-1. Representation of the point frame sampling procedure for selecting NSS Stage I reaches .. 53
5-2. The pH-ANC relationship for samples with ANC < 200 peq/L taken at the downstream
nodes of stream reaches in the NSS in subregions 1D, 2B, and 2C 61
5-3. Definition of soil sampling classes for the DDRP Soil Survey in the M-APP Region 76
5-4. Selection of starting points for sampling 79
5-5. Field selection of a sampling point for sampling class on a watershed 80
5-6. Quality assurance and quality control soil samples for mineral and organic batches
in the M-APP Region 97
5-7. Procedure for calculating regional area in each sampling class 105
5-8. Relative areas of sampling classes in soils of the DDRP target population in the
M-APP Region 106
5-9. Aggregated soil variables for individual pedons in the M-APP Region 107
5-10. Procedure for calculation of cumulative distribution function for a soil variable
in the M-APP Region 109
5-11. Cumulative distribution functions for pedon-aggregated soil variables for the M-APP
Region 110
5-12. Sulfur deposition scenarios for the M-APP Region for Level II and III analyses 114
5-13. Example of average annual runoff map for 1951-80 128
VIII
-------
7-1. Percent sulfur retention for intensively studied sites in the United States and
Canada relative to the southern extent of the Wisconsinan glaciatibn 141
7-2. Population-weighted distribution of projected percent sulfur retention, with upper
and lower bounds for 90 percent confidence intervals, for NSWS target population
of (A) lakes in the Northeast, and for stream reaches in (B) the Mid-Appalachian,
and (C) Southern Blue Ridge Province Regions 145
7-3. Population-weighted distributions of projected percent sulfur retention with upper
and lower bounds for 90 percent confidence intervals, for NSS target population
systems in four additional NSS subregions: (A) Southern Appalachian Plateau,
(B) Mid-Atlantic Coastal Plain, (C) Catskills/Poconos, and (D) Piedmont 148
7-4. Comparisons of regional distributions of (A) measured sulfate concentration, (B) steady-
state sulfate concentration, (C) percent suifur retention and (D) change in sulfate
concentration to steady state for DDRP target populations of lakes or streams in the
NE, M-APP, and SBRP 149
9-1. Deposition scenarios used for modelling future sulfate dynamics in soils of DDRP M-APP
watersheds 164
9-2. Equilibrium soil solution sulfate (B horizon) and measured stream sulfate concentration
for the 36 DDRP M-APP watersheds . 172
9-3. Measured and modelled sulfate concentrations and percent sulfur retention for the
36 DDRP M-APP watersheds 173
9-4. Distributions of measured and modelled sulfate concentration and percent sulfur
retention for the M-APP target population of 5,496 stream reaches 176
9-5. Distributions of measured sulfate concentration, steady state sulfate concentration,
percent sulfur retention, and change in sulfate concentration to sulfur steady state
in the NE, M-APP, and SBRP 178
9-6. Time trends for modelled inputs and outputs of sulfate for soils on a typical
watershed in the M-APP Region, and for the fastest and slowest responding systems
in the region 181
9-7. Projected changes in sulfate concentration, delta sulfate (change from base year
concentration), and percent sulfur retention for M-APP DDRP target stream reaches 183
9-8. Distribution of time to steady state for sulfur for soils on DDRP target stream reach
watersheds in the M-APP Region 184
9-9. Distributions of time to steady state for sulfur in soils on DDRP target watersheds
in the NE, M-APP, and SBRP 187
9-10. Projected temporal trends for sulfate in a typical M-APP watershed, and for the
fastest and slowest responding systems in the region, for three scenarios of future
sulfur deposition 188
9-11. Extent and duration of lags in modelled sulfate response for soils on 36 M-APP
watersheds following a 50 percent decrease in sulfur deposition 190
9-12. Flow diagram for the one-box Bloom-Grigal soil simulation model 200
IX
-------
9-13. Regional CDFs of Bloom-Grigal model projected base saturations of soils on the
DDRP target population of M-APP stream watersheds for three deposition scenarios
(A, B, and C) at 0, 20, 50, and 100 years 211
9-14. Regional CDFs of Bloom-Grigal model projected pH of soils on the DDRP target
population of M-APP stream watersheds for three deposition scenarios (A, B, and C)
at 0, 20, 50, and 100 years, using deposition data for the full complement of base
cations in dry deposition 216
9-15. The curvilinear relationship between soil pH and base saturation for typical values of n
and pKa as defined by Equation 9-6 220
9-16. Regional CDFs of Bloom-Grigal model projected soil solution concentrations of
AI3+ of so:ls on the DDRP target population of M-APP stream watersheds for
three deposition scenarios (A, B, and C) at 0, 20, 50, and 100 years 222
9-17. Projected changes in median base saturation for 100-year simulations, for soils
in DDRP target population watersheds in the NE, M-APP, and SBRP regions,
generated using the Bloom-Grigal model for a scenario of constant deposition at
current levels 228
9-18. Projected changes in median pH, for 100-year simulations, for soils in DDRP
target populations in the NE, M-APP, and SBRP regions, generated using the
Bloom-Grigal model for a scenario of constant deposition at current levels 229
9-19. Projected changes in median soil solution AI3+, for 100-year simulations, for soils
in DDRP target populations in the NE, M-APP, and SBRP regions, generated using the
Bloom-Grigal model for scenarios of constant deposition at current levels 230
9-20. Hydrologic pathways and soil geochemical processes considered by the modified Reuss
model used for DDRP simulations 233
9-21. Cumulative distribution of year 0 calibrated values of ANC, for the target population
of DDRP watersheds in the M-APP Region, based on simulations using the Reuss model
with Typical Year deposition estimates 238
9-22. Comparison of measured values of ANC for DDRP watersheds in the M-APP Region
and values simulated using the modified Reuss model with Typical Year deposition
estimates 240
9-23. Cumulative distributions of projected changes in ANC at 20, 50, and 100 years,
for the target population of DDRP M-APP watersheds, based on Reuss model
simulations with Typical Year deposition estimates 242
10-1. Three deposition scenarios used in the M-APP Region 258
10-2. Comparison for consistency among different USGS gages within the M-APP Region
for monthly flow fractions 260
10-3. Simulated and observed values for ANC and flow in Coweeta watershed 36 using the
MAGIC model 269
10-4. Simulated and observed values for calcium and sulfate concentrations in Coweeta
watershed 36 using the MAGIC model 270
-------
10-5. Simulated and observed values for ANC and flow in Coweeta watershed 34 using
the MAGIC model 272
10-6. Simulated and observed values for calcium and sutfate concentrations in Coweeta
watershed 34 using the MAGIC model 273
10-7. MAGIC calibrated values at year 0 and NSS Phase I values for ANC, sulfate,
calcium plus magnesium, and pH for the DDRP M-APP sample watersheds 277
10-8. MAGIC calibrated values at year 0 and NSS Phase I values for ANC, calcium plus
magnesium, pH, and sulfate for the DDRP M-APP target population 278
10-9. Projections of ANC for the M-APP stream target population under current sulfur
deposition (Scenario A), 50 percent deposition decrease (Scenario B), and 50/20
percent deposition decrease (Scenario C) over 100 years 282
10-10. Projections of pH for the M-APP stream target population under current sulfur
deposition (Scenario A), 50 percent deposition decrease (Scenario B), and 50/20
percent deposition decrease (Scenario C) over 100 years 286
10-11. Projections of sulfate concentration for the M-APP stream target population under current
sulfur deposition (Scenario A), 50 percent deposition decrease (Scenario B), and 50/20
percent deposition decrease (Scenario G) over 100 years 288
10-12. Projections of calcium plus magnesium for the M-APP stream target population under
current sulfur deposition (Scenario A), 50 percent deposition decrease (Scenario B),
and 50/20 percent deposition decrease (Scenario C) over 100 years 289
10-13. Box and whisker plots of projected ANC for M-APP streams under current deposition
(Scenario A) 294
10-14. Box and whisker plots of projected ANC for M-APP streams under a 50 percent
deposition decrease (Scenario B) 295
10-15. Box and whisker plots of projected ANC for M-APP streams under a 50/20 percent
deposition decrease (Scenario C) 296
10-16. Box and whisker plots of projected pH for M-APP streams under current deposition
(Scenario A) 297
10-17. Box and whisker plots of projected pH for M-APP streams under a 50 percent deposition
decrease (Scenario B) 298
10-18. Box and whisker plots of projected pH for M-APP streams under a 50/20 percent
deposition decrease (Scenario C) 299
10-19. Projected change in median ANC concentration over 50 years for all three deposition
scenarios as a function of the NSS Phase I ANC used to classify DDRP streams in the
target population 301
10-20. Change in median pH concentration over 50 years for all three deposition scenarios
as a function of the NSS Phase I ANC used to classify DDRP streams in the
target population 303
XI
-------
10-21. Change in pH for sampled streams in the DDRP M-APP Region as a function of the
initial calibrated pH at year 0 304
11-1. Percentage of stream reaches of the DDRP M-APP modelling target population projected
by the MAGIC model to have ANC < 0 ^eq/L at the lower reach nodes 328
11-2. Length of stream reaches of the DDRP M-APP modelling projected by the MAGIC
model to have ANC < 0 fteq/L at the lower reach nodeso 329
11-3. Percentage of stream reaches of the DDRP M-APP modelling projected by the MAGIC
model to have ANC < 50 fteq/L at the lower reach nodes 330
11-4. Length of stream reaches of the DDRP M-APP modelling projected by the MAGIC
model to have ANC < 50 ^weq/L at the lower reach nodes 331
XII
-------
1-1.
1-2.
1-3.
1-4.
1-5.
1-6.
1-7.
2-1.
5-1.
5-2.
5-3.
5-4.
7-1.
PLATES
Direct/Delayed Response Project study regions and sites
PAGE
. . 4
Sulfur retention and wet sulfate deposition for National Surface Water Survey
subregions in the eastern United States
11
Changes in sulfur retention up to 100 years for watersheds of the M-APP Region as
projected by the MAGIC modeller its specified DDRP target population 12
Changes in ANC up to 100 years for stream reaches of the M-APP Region as projected by
the MAGIC model for its specified DDRP target population 16
Changes in pH up to 100 years for stream reaches of the M-APP Region as projected by
the MAGIC model for its specified DDRP target population 17
Changes in median ANC at 50 years for stream reaches of the M-APP Region as projected
by the MAGIC model for its specified DDRP target population 18
Changes in median pH at 50 years for stream reaches of the M-APP Region as projected
by the MAGIC model for its specified DDRP target population 19
Direct/Delayed Response Project study regions and sites 35
ANC at lower nodes of DDRP stream reaches in the M-APP Region 58
ADS and NCDC sites linked with DDRP study sites for the M-APP Region . 117
Pattern of typical year sulfate deposition for the DDRP M-APP study sites 125
Pattern of typical year sulfate deposition for all DDRP study sites 126
Sulfur retention and wet sulfate deposition for National Surface Water Survey
subregions in the eastern United States 146
11-1. Sulfur retention and wet sulfate deposition for National Surface Water Survey
subregions in the eastern United States 318
11-2. Changes in sulfur retention up to 100 years for watersheds of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population 320
11-3. Changes in ANC up to 100 years for stream reaches of the M-APP Region as projected
by the MAGIC model for its specified DDRP target population 323
11-4. Changes in pH up to 100 years for stream reaches of the M-APP Region as projected
by the MAGIC model for its specified DDRP target population . 324
11-5. Changes in median ANC at 50 years for stream reaches of the M-APP Region as projected
by the MAGIC model for its specified DDRP target population 325
11-6. Changes in median pH at 50 years for stream reaches of the M-APP Region as projected
by the MAGIC model for its specified DDRP target population 326
XIII
-------
TABLES
5-1.
5-2.
5-3.
5-4.
5-5.
5-6.
5-7.
5-8.
5-9.
5-10.
5-11.
5-12.
5-13.
5-14.
5-15.
5-16.
7-1.
7-2.
8-1.
PAGE
Stream Identification (ID), Weight, and Inclusion Probabilities for the M-APP DDRP
Sample Watersheds 55
Stream Identification (ID) and Name, and State and Latitudinal/Longitudinal Location
of the M-APP Sample Watersheds, Sorted by Stream ID 56
Stream Identification (ID) and Name, Sorted by State - M-APP Sample Watersheds 57
Population Estimates of Combined Number, Length, and Percentages of Stream Reaches
Having ANC Less Than Selected Reference Values at Downstream Nodes During Spring
Baseflowfor NSS Phase I and DDRP M-APP Region 63
List of DDRP Watersheds by Geographic Group 82
Laboratory Analysis of DDRP Soil Samples 85
Analytical Variables Measured in the DDRP Soil Survey 87
Within-Batch Precision Objectives for the Analytical Measurement Quality Objectives 92
Contract-Required Detection Limits for the Analytical Laboratories 94
Detection Limits for Evaluation of Contractual Compliance and for Independent
Assessment of Analytical and System-Wide Measurement 96
Attainment of Measurement Quality Objectives for Precision by the Analytical
Laboratories as Determined from Blind Audit Samples for the M-APP Region 98
Analytical Bias Estimates for Mineral Samples -101
Soil Chemistry Relationships and Delimiters Used by LEVIS 103
Medians of Pedon-Aggregated Values of Soil Variables for the M-APP Region 112
Monthly Values of Leaf Area Index (LAI) Used to Apportion Annual Dry Deposition to
Monthly Values 119
Ratios of Dry Deposition to Wet Deposition for DDRP Study Sites for the Typical
Year (TY) Deposition Datasets 122
Percent Sulfur Retention - Summary Statistics by Region for Lakes and Stream
Reaches in the NSWS (ELS) and NSS 144
Summary Statistics for Distributions of Sulfate Concentration, Steady-State Sulfate
Concentration, Change in Sulfate Concentration to Steady State, Percent Sulfur
Retention for DDRP Target Population Watersheds in the NE, M-APP, and SBRP 150
Summary of Level I Regression Analyses for DDRP Watersheds in the NE and SBRP,
Listing Values of R2 for Analyses Using Different Groups of Watershed Attributes . . .' 155
XIV
-------
9-1. Summary Comparison of Adsorbed Sulfate and Adsorption Capacities of Soils on
Watersheds in the DDRP Target Populations in the M-APP, NE, and SBRP 170
9-2. Rates of Increase for Sulfate in Stream Systems in the M-APP Region, Comparing
Model Projections to Measured Rates of Increase 175
9-3. Summary of Sulfur Status in Surface Waters for Watersheds in DDRP Target Populations
in the NE, M-APP, and SBRP 177
9-4. Comparison of Sulfate Concentrations and Sulfur Budget Status for DDRP Watersheds in
Subregions of the M-APP Region 179
f t
9-5. Summary Statistics for Modelled Changes in Sulfate Concentration and Percent
Sulfur Retention, and for Changes in Sulfate Concentration for Watersheds in the
M-APP Region 185
9-6. Summary Comparison of Modelled Changes in Sulfate Concentration in M-APP Stream
Systems Under Three Future Sulfur Deposition Scenarios 191
9-7. List of Input Data for the Bloom-Grigal Soil Acidification Model 205
9-8. Regionally-Weighted Median Values of Base Year Annual Acid Inputs for the
Bloom-Grigal Model for Watersheds in the M-APP Regiqn, for Three Levels of Dry
Base Cation Deposition 207
9-9. Regionally-Weighted Median Values of Annual Initial Soil Chemical Values of the Level II
Bloom-Grigal Modelling for the M-APP, NE, and SBRP 209
9-10. Bloom-Grigal Model Regional Projections of Percent Base Saturation for Soils in the
M-APP Region as a Function of Time and Deposition Scenario 212
9-11. Bloom-Grigal Model Regional Projections of Percent Base Saturation for Soils in the
M-APP Region as a Function of Time and Deposition Scenario and for a Scenario of a
50 Percent Reduction in Base Cation Dry Deposition 213
9-12. Bloom-Grigal Model Regional Projections of Percent Base Saturation for Soils in the
M-APP Region as a Function of Time and Deposition and for a Scenario of No Base
Cations in Dry Deposition 214
9-13. Bloom-Grigal Model Regional Projections of Soil Solution pH for Soils in the
M-APP Region as a Function of Time and Deposition Scenario for a Scenario of a Full
Complement of Dry Base cations in Dry Deposition 217
9-14. Bloom-Grigal Model Regional Projections of Soil Solution pH for Soils in the
M-APP Region as a Function of Time and Deposition Scenario and for a Scenario of a 50
Percent Reduction in Base Cations in Dry Deposition 218
9-15. Bloom-Grigal Model Regional Projections of Soil Solution pH for Soils in the M-APP
Region as a Function of Time and Deposition Scenario and for a Scenario of Zero
Base Cations in Dry Deposition 219
9-16. Bloom-Grigal Model Regional Projections of Soil Solution AI3+ for Soils in the M-APP
Region as a Function of Time and Deposition Scenario for the Scenario of a Full
Complement of Base Cations in Dry Deposition 223
xv
-------
9-17. BIoom-Grigal Model Regional Projections of Soil Solution AI3+ for Soils in the M-APP
Region as a Function of Time and Deposition Scenario for the Scenario of a 50 Percent
Reduction of Base Cations Concentration in Dry Deposition 224
9-18. Bloom-Grigal Model Regional Projections of Soil Solution AI3+ for Soils in the M-APP
Region as a Function of Time and Deposition Scenario for the Scenario of Zero
Base Cations in Dry Deposition 225
9-19. Summary Statistics for Regional Medians for Base Saturation (BS), pH, and Soil
Solution AI3+ Concentration for Soils on DDRP Target Populations of Watersheds
in the NE, M-APP, and SBRP 231
9-20. Summary Statistics for Current ANC in DDRP Target Population Stream Reaches in the
M-APP Region 239
9-21. Summary Statistics for Projected Changes in Surface Water ANC Values in Streams
in the DDRP Target Population in the M-APP Region at 20, 50, and 100 Years
for Three Deposition Scenarios 243
9-22. Summary Statistics for Projections of Current ANC, Predicted Using the Reuss Model
with Typical Year Deposition Estimates for the NE, M-APP and SBRP 246
9-23. Projected Changes in ANC at 20, 50, and 100 Years, Based on Reuss Model Analyses
Using Typical Year Deposition Estimates for the NE, M-APP, and SBRP 247
10-1. Level III Operational Assumptions 256
10-2. Monthly Runoff Fractions for Selected USGS Gaging Stations 261
10-3. Inputs, Watershed Characteristics, and Surface Water Chemistries for Coweeta
Watersheds 266
10-4. RMSE Values Based on Volume-Weighted Monthly Averages of Simulated and Observed
Quantities for Coweeta Watersheds over a 12-Year Period of Record 268
10-5. Watersheds and Attributes for Which Satisfactory MAGIC Calibration Was
Not Achieved 275
10-6. Descriptive Statistics of Calibrated ANC, pH, Sulfate Concentration, and Calcium
Plus Magnesium Using MAGIC Compared with NSS Observed Values 279
10-7. Population Estimates of Combined Number, Length and Percentages of Stream
Reaches Having ANCs Less than Selected Reference Values at Downstream Nodes
During Spring Baseflow for NSS Phase I; All M-APP DDRP; DDRP Systems Simulated . . 280
10-8. Descriptive Statistics of Projected ANC, pH, Sulfate, Percent Sulfur Retention,
and Calcium and Magnesium Using MAGIC for Current and Decreased Deposition 283
10-9. Change Projected by the MAGIC Model in the Number of M-APP DDRP Lower
Reach Nodes Having (A) ANC < 0 or (B) ANC < 50 ^eq/L 290
10-10.
Percentage of M-APP Target Population Stream Reaches Projected by the
MAGIC Model to Have ANC (A) < 0 or (B) ANC < SO^eq/L at the Lower Reach Nodes
2S1
XVI
-------
10-11. Length of M-APP Target Population Stream Reaches Projected by the MAGIC Model
to Have (A) ANC < 0 or (B) ANC < 50 /*eq/L at the Lower Reach Nodes 292
10-12. Changes in Median ANC, pH, Sulfate Concentration, and Calcium plus Magnesium
Concentration over 50 and 100-Year Periods as a Function of the Initial NSS
ANC Groups and Deposition Scenarios 300
10-13. Effects of Critical Assumptions and Uncertainties on Projected Rates of Change 307
10-14. Comparison of ANC Values at Upstream and Downstream Reach Nodes for the M-APP
DDRP Modelling Target Population 314
xvn
-------
PRIMARY CONTRIBUTORS
The Direct/Delayed Response Project and this Final Draft Report represent the efforts of many
scientific, technical, and support staff. The primary contributors to this report are noted here.
Section 1: Executive Summary
M.R. Church, U.S. Environmental Protection Agency
P.W. Shaffer, ManTech Environmental Technology, Inc.
Section 2: Introduction
M.R. Church, U.S. Environmental Protection Agency
Section 3: Processes of Acidification
P.W. Shaffer, ManTech Environmental Technology, Inc.
Section 4: Project Approach
M.R. Church, U.S. Environmental Protection Agency
Section 5: Data Sources and Descriptions1
D.L Cassell, IManTech Environmental Technology, Inc.
M.R. Church, U.S. Environmental Protection Agency
D.A. Lammers, U.S.D.A. Forest Service
JJ. Lee, U.S. Environmental Protection Agency
Others Providing Support Contributions
LJ. Blume, U.S. Environmental Protection Agency
C.C. Brandt, Oak Ridge National Laboratory
G.E. Byers, Lockheed Engineering and Sciences Co.
J.S. Kern, ManTech Environmental Technology, Inc.
LH. Liegel, U.S.D.A. Forest Service
D.C. Mortenson, ManTech Environmental Technology, Inc.
CJ. Palmer, ManTech Environmental Technology, Inc.
M.L. Papp, Lockheed Engineering and Sciences Co.
S.M. Pierson, ManTech Environmental Technology, Inc.
B.P. Rochelle, ManTech Environmental Technology, Inc.
D.D. Schrnoyer, Martin Marietta Energy Systems, Inc.
BA Schumacher, Lockheed Engineering and Sciences Co.
R.S. Turner, Oak Ridge National Laboratory
Section 6: Regionalization of Analytical Results
D.L. Cassell, ManTech Environmental Technology, Inc.
D.L. Stevens, Eastern Oregon State University
Sectiorf 7: Watershed Sulfur Retention
P.W. Shaffer, ManTech Environmental Technology, Inc.
M.R. Church, U.S. Environmental Protection Agency
B.P. Rochelle, ManTech Environmental Technology, Inc.
Section 8: Level I Statistical Analyses
P.W. Shaffer, ManTech Environmental Technology, Inc.
1 Contributors to this section listed alphabetically
xviii
-------
Section 9: Level II Single-Factor Time Estimates
P.W. Shaffer, ManTech Environmental Technology, Inc.
M.G. Johnson, ManTech Environmental Technology, Inc.
G.R. Holdren, ManTech Environmental Technology, Inc.
Section 10: Level III Dynamic Watershed Models
K.W. Thornton, FTN & Associates, Ltd.
M.R. Church, U.S. Environmental Protection Agency
C.I. Liff, Utah State University
P.W. Shaffer, ManTech Environmental Technology, Inc.
Extramural Cooperator Providing Modelling Expertise and Support:
B.J. Cosby, University of Virginia
Section 11: Summary of Results
M.R. Church, U.S. Environmental Protection Agency
P.W. Shaffer, ManTech Environmental Technology, Inc.
Section 12: References
Section 13: Glossary
Graphics:
Color
S.M. Pierson, ManTech Environmental Technology, Inc.
G.D. Bishop, ManTech Environmental Technology, Inc.
Black and White:
J. Berglund, Albany, OR
T. Thompson, FTN & Associates, Ltd.
Editing:
S.J. Christie, ManTech Environmental Technology, Inc.
Word Processing:
C.B. Roberts, ManTech Environmental Technology, Inc.
xix
-------
ACKNOWLEDGMENTS
The completion of the Direct/Delayed Response Project (DDRP) and the preparation of this report
have required the efforts of many scientists and support personnel. We acknowledge here a few of
those persons who made particularly outstanding contributions. To ail the others who helped us, but
who are not named here, we also extend our sincere thanks. We also thank again (but do not
necessarily list again) all those who assisted in the development of the Project and the production of
the first major results report (Church et al., 1989).
We thank Courtney Riordan of the EPA Office of Research and Development (ORD), Rick
Linthurst, Dan McKenzie and Dixon Landers (Directors of the Aquatic Effects Research Program) and
Tom Murphy, Laboratory Director for EPA's Environmental Research Laboratory-Corvallis (ERL-C), for
their continued support and interest in the Project.
Dennis Leaf and Brian McLean from the Office of Air and Radiation at EPA-Headquarters provided
insight and suggestions for analyses of particular relevance to questions of Agency policy. We thank
them for their interest and assistance.
Paul Ringold (ORD), Bill Fallpn (ORD), and Chuck Frank (ERL-C) and his staff, all have provided
much administrative assistance to the Project. We thank them all for their efforts and assistance.
Many landowners and state and government agencies allowed us to map and sample soils on
their properties. We thank them for their permission to do so.
The cooperation of the U.S. Department of Agriculture (USDA) Soil Conservation Service (SCS)
was absolutely essential to the completion of the DDRP Soil Survey. People in the SCS state offices
who were responsible for mapping of DDRP watersheds and obtaining the soil descriptions included
Garland Lipscomb and George Martin (Pennsylvania) and William Hatfield and Cameron Loerch (West
Virginia). We thank them as well as the other approximately 20 soil scientists from these states who
played such an important part in the Soil Survey. In addition, we thank Milton Meyer and Dick Arnold
of the National Office, and Henry Mount of the National Technical Centers of the SCS for their advice
and able assistance.
John Warner was the Regional Correlator/Coordinator for the soil mapping and sampling. We
thank him for his continued work with the DDRP and the excellent job that he did.
A large and dedicated staff at EPA's Environmental Monitoring and Systems Laboratory-Las
Vegas (EMSL-LV) played a crucial role in support of the DDRP soil survey. Gareth Pearson and Bob
Schonbrod provided supervisory guidance for the DDRP Soil Survey activities at EMSL-LV. Lou Blume
was responsible for contracting and management of the soil preparation and analytical laboratories,
and for the delivery of operations reports, quality assurance reports, and methods manuals for the
DDRP. Anders Denson (EMSL-LV) and Marian Bernd (EPA-RTP) provided critical contract
administration support. Mike Papp of Lockheed Engineering and Sciences Corporation (LESC)
supervised the delivery of verified field, soil preparation and analytical databases for the DDRP. Rick
xx
-------
Van Remortel (LESC) assisted in the verification of the field, preparation, and analytical databases, and
corresponding reports. He also supervised the soil preparation laboratory operations. Brian
Schumacher (LESC) was the technical lead for verification of the analytical database and assisted in
the development of the analytical methods as well as the QA reports. Rod Slagle (LESC) served as
the DDRP soils database manager at EMSL-LV. Mohammed Miah (LESC) and John Teberg (LESC)
provided statistical and data analysis support. Craig Palmer and Barbara Conkling (Environmental
Research Center, University of Nevada Las Vegas) provided invaluable technical assistance for various
quality assurance aspects.
Three contract laboratories provided valuable analytical support in the form of (1) high quality
soils analyses, and (2) continual assistance to the quality assurance staff in interpreting the effects and
implications of analytical approaches. The laboratories and the key people were: Weyerhaeuser
Analytical Services, Kari Doxsee; Harris Environmental Technologies, Tom Buelt; and Huffman
Laboratories, Ed Huffman. We thank them for all of their hard work.
Deborah Coffey of ManTech Environmental Technology, Inc. (METI) played a critical role in
ensuring the quality of the watershed and soils data gathered for the Project. She either had a major
responsibility for, or assisted in, the development of data quality objectives, field sampling manuals,
laboratory methods manuals, field operations reports, field quality assurance reports and numerous
other facets of the Soil Survey. We thank her for her unswerving attention to detail.
Numerous extramural cooperators assisted in this Project. Warren Gebert, Bill Krug, David
Graczyk and Greg Allord of the U.S. Geological Survey (Madison, Wisconsin) supplied runoff data and
maps that were crucial to the Project. Wayne Swank and Jack Waide of the USDA Forest Service
cooperated with the Project in allowing us to use data gathered by the Coweeta Hydrologic
Laboratory. Jack Waide also provided many insights into the wbrkings of watersheds in the Southern
Blue Ridge and in the application of watershed simulation models. Tony Olsen, Sally Wampler and
Jeanne Simpson of Battelle Pacific Northwest Laboratories provided a great deal of information on
estimates of wet deposition to sites of interest in the Eastern United States. Tony Olsen also assisted
in editing text describing analyses of the wet deposition data. Robin Dennis and Terry Clark of the
EPA's Atmospheric and Exposure Assessment Laboratory-Research Triangle Park and Steve Seilkop
of Analytical Services, Incorporated, provided key information on estimates of atmospheric dry
deposition. Steve Lindberg of Oak Ridge National Laboratory and Bruce Hicks and Tilden Myers of
the National Oceanographic and Atmospheric Administration provided considerable assistance in the
form of discussions and preliminary data on rates of atmospheric dry deposition. We thank all of
these cooperators for their assistance.
The accomplishments of any project are due in a large measure to assistance from peer
reviewers. The DDRP benefited immensely from peer review comments from beginning to end.
The following scientists served as reviewers of the initial Review Draft Report: Jim Wigington of
ERL-C, Rick Webb of the University of Virginia, Mike Bowman of the Maryland Department of Natural
Resources, Scott Overton of Oregon State University, David Grigal of the University of Minnesota, Phil
Kaufmann and Alan Herlihy of Utah State University, John Melack of the University of California - Santa
XXI
-------
Barbara, and Pat Ryan of Oak Ridge National Laboratory. This report benefited greatly from their
comments and constructive criticisms.
Numerous other scientists also served as reviewers over the years of individual aspects of the
Project or of the Project as a whole. We thank them for helping us to improve the quality of our work.
John Berglund of InstaGraphics, Inc. prepared many of the figures that appear in this report.
We thank him for his skillful efforts.
Susan Christie (METI) edited this report and did her best to see that it made sense and read
well. Any success that we have achieved along these lines is due mainly to her painstaking efforts.
All of the word processing for this report was done by Carol Roberts (METI). We thank Carol for
her many, many hours of hard work and for her valuable assistance in producing this report.
Once again the DDRP Technical Director sincerely thanks all of the Project staff, associated
contract scientists and support personnel, and extramural cooperators for their contributions to this
work. It has been a memorable effort. Thank you.
XXII
-------
PREFACE
This report describes the investigation of potential long-term effects of atmospheric deposition of
sulfur on stream water chemistry in the Mid-Appalachian Region (M-APP) of the United States. These
analyses have been pursued as part of the U.S. EPA's Direct/Delayed Response Project (DDRP)
(Church, 1989). Previous analyses of potential effects on lakes in the Northeast (NE) and stream
reaches in the Southern Blue Ridge Province (SBRP) were reported by Church et al. (1989).
The analyses presented here parallel those for the NE and SBRP with two notable exceptions.
First, the extensive statistical analyses performed on relationships among atmospheric deposition and
soil and surface water characteristics for the NE and SBRP were not repeated for the M-APP. Those
previous analyses indicated that processes considered within the DDRP soil and watershed modelling
work appeared to account reasonably well for extant relationships. Repetition of such extremely
time-consuming analyses for the M-APP would not have been an economical or especially fruitful
endeavor: The second notable difference between analyses for the M-APP and those for the NE and
SBRP was in the dynamic watershed modelling. For the NE and SBRP, we used three independent
dynamic watershed models: (1) the Integrated Lake-Watershed Acidification Study (ILWAS) model,
(2) the Enhanced Trickle-Down Model (ETD), and (3) the Model of Acidification of Groundwater in
Catchments (MAGIC). This model intercomparison work led to three important conclusions: (1) the
work was extremely time consuming and expensive, (2) the MAGIC model was the most readily
calibrated and exercised of the three models, and (3) the projections of future effects on a regional
scale (the thrust of the DDRP) provided by the models were reasonably similar. Therefore, we decided
to apply only the MAGIC model to the stream reaches studied in the Mid-Appalachian Region.
We present the results of our work for the M-APP in a format as nearly as..possible like that of
our previous reporting for the NE and SBRP (Church et al., 1989). Our intent is to make it easy to
compare results among all of the regions studied in the DDRP. In this report, we also make
interregional comparisons of results where appropriate. In some cases, we repeat (often at a reduced
level) introductory material from our previous report. We do this to assure a reasonable level of
self-containment or completeness in this single report for the Mid-Appalachian Region. We would like
for it to "stand alone" as well as possible. Those readers who are familiar with our previous report can
easily progress to pertinent reporting of results; those readers without access to or less familiar with
our report for the NE and SBRP will, hopefully, find a more than adequate description of our work
effort here.
xxiii
-------
PREFACE TO THE EXECUTIVE SUMMARY
This Executive Summary contains both a summary of Project analyses and an overview of the
Project background and approach. Those readers wishing a synopsis only of major Project results
may turn directly to Section 1.4. Because of the complex design and approach of this Project,
however, we encourage readers to review Sections 1.1 through 1.3 of this Executive Summary.
M. Bobbins Church, Technical Director
Direct/Delayed Response Project
XXIV
-------
SECTION 1
EXECUTIVE SUMMARY
1.1 INTRODUCTION
1.1.1 Project Background
Much scientific interest and public debate surround the subject of the effects of acidic deposition
on freshwater ecosystems (e.g., Schindler, 1988; Mohnen, 1988). The U.S. Environmental Protection
Agency (EPA) recently completed a comprehensive chemical survey, the National Surface Water
Survey (NSWS), of lakes and streams in the United States considered to be most vulnerable to acidic
deposition (i.e., those with the lowest acid neutralizing capacity, or ANC) (Linthurst et al., 1986;
Landers et al., 1987; Kaufmann et al., 1988). Analysis of these and other lake and stream chemistry
data, together with data on temporal and spatial patterns of atmospheric deposition, indicates that
long-term deposition of sulfur-containing compounds originating from the combustion of fossil fuels
has acidified (i.e., decreased the ANC of) some surface waters in eastern North America (Altshuller
and Linthurst, 1984; NAS, 1986; Sullivan et al., 1988; Neary and Dillon, 1988; Asbury et al., 1989;
Baker et al., 1991).
The transport of mobile anions (primarily sulfate) through watershed soils and closely associated
cation leaching are the most widely accepted mechanisms of this acidification process (Seip, 1980;
Galloway et al., 1983; Driscoll and Newton, 1985; Church and Turner, 1986). Krug (see Krug and
Frink, 1983; Krug et al., 1985; Krug, 1989) has hypothesized that acidic deposition may have shifted
the nature of some very low ANC or naturally acidic surface waters from organic acid "dominance" to
mineral acid dominance. A number of studies have examined or have shed light upon this question
from the perspective of soil processes (e.g., Krug and Isaacson, 1984; McColl and Pohlman, 1986),
observations of surface water chemistry (Driscoll et al., 1988; Driscoll et al., 1989), or manipulation
experiments (Hedin et al., 1990), but have not confirmed any widespread importance of this phenom-
enon. Examination of historical changes in the surface water chemistry of Adirondack lakes (as
reconstructed from patterns of changes in diatom and chrysophyte populations), in fact, show no con-
sistent decreases in DOC (Gumming et al., In press) and argue against the hypothesis. In any event,
the net effect of atmospheric deposition of sulfuric acid on surface water chemistry is a shift toward
aquatic systems more dominated by mineral acidity and more likely to contain inorganic forms of
aluminum, which are toxic to aquatic organisms.
Given that acidification of some surface waters has occurred, critical scientific and policy ques-
tions focus on whether acidification is continuing in the regions of concern, whether it is just beginning
in other regions, how extensive the effects may become, and over what time scales the effects may
occur. The EPA is examining these questions through the activities of the Direct/Delayed Response
Project (DDRP) (Church and Turner, 1986; Church, 1989). The Project began in 1984 at the specific
request of the EPA Administrator, following a meeting of the Panel on Processes of Lake Acidification
of the National Academy of Sciences (NAS). Principal among the conclusions of the Panel was that
atmospheric deposition of sulfur-containing compounds is the major source of long-term surface water
acidification in eastern North America (NAS, 1984). The Panel also debated at length the dynamic
aspects of the acidification process. The DDRP was designed to focus on the topic of acidification
-------
dynamics and draws its name from consideration of whether acidification is immediate (immediately
proportional to levels of deposition, i.e., "direct") or whether it lags in time (i.e., is "delayed") because of
edaphic characteristics. Church and Turner (1986) presented discussion of the processes of long-
term surface water acidification and methods for its investigation at the beginning of the Project. Major
points of that discussion are repeated in Section 3 of this report. Church et al. (1989) presented a
relatively brief and more current discussion of processes especially relevant to this Project. Even more
recently, Turner et al. (1990) have thoroughly reviewed processes that both control and are associated
with long-term surface water acidification.
Although recent research has indicated that deposition of nitrogen-containing compounds is
potentially important to both episodic (Galloway et al., 1987; Driscoll et al., 1987; Wigington et al.,
1990) and long-term (Henriksen and Brakke, 1988) acidification of surface waters, the DDRP does not
address these effects. Such effects are the focus of developing or ongoing research within other EPA
research programs.
1.1.2 Primary Objectives
The DDRP has four major technical objectives related to atmospheric/terrestrial/aquatic inter-
actions:
• to describe the regional variability of soil and watershed characteristics,
• to determine which soil and watershed characteristics are most strongly related to surface
water chemistry,
• to estimate the relative importance of key watershed processes in moderating regional effects
of acidic deposition, and
• to classify a sample of watersheds with regard to their response to inputs of acidic deposition
and to extrapolate the results from this sample of watersheds to the study regions.
The fourth objective is the critical "bottom line" of the Project.
It was never the intent of the DDRP to serve as a research project on the exact mechanisms and
processes of surface water acidification. Rather, the principal mandate of the Project was to make
regional projections of the future effects of sulfur deposition on long-term surface water chemistry,
based on the best available data and most widely accepted hypotheses of the acidification process.
In-depth investigations into processes of soil and surface water acidification are being conducted by
other projects within the National Acid Precipitation Assessment Program (NAPAP).
1.1.3 Study Regions
The Project focuses on three regions of the eastern United States where low ANC surface waters
are located and where levels of atmospheric deposition are greatest (relative to other U.S. regions):
(1) the Northeast (NE), (2) upland areas of the Mid-Atlantic (referred to here as the Mid-Appalachian
-------
Region), and (3) the mountainous section of the Southeast called the Southern Blue Ridge Province
(SBRP) (Plate 1-1). Initiation of the Project depended on the availability of the regional surface water
chemistry data of the NSWS. Thus, the Project focused its initial work on the lake resources of the NE
(Linthurst et al., 1986) and the stream resources of the SBRP (Messer et al., 1986). Church et al.
(1989) presented the results for those two regions. Recent stream survey work by Kaufmann et al.
(1988) has allowed investigation of possible future effects of acidic deposition on stream chemistry in
the Mid-Appalachian (M-APP) Region. This report presents the results of our work in this region.
1.1.4 Time Frames of Concern
The DDRP focuses on the potential effects of acidic deposition on surface water ANC at key
annual "index" periods, as defined by the NSWS sampling periods (i.e., the fall turnover period of
complete mixing for lakes and the spring baseflow period for streams; see Section 5.3). The primary
time horizon for DDRP analyses is 50 years, which was selected on the basis of the projected lifetimes
of existing power plants and the potential implementation of additional emission controls relative to
those lifetimes. Where possible and reasonable, some time-dependent analyses are extended beyond
this 50-year horizon to better evaluate process rates and changes and potential effects.
1.2 PROCESSES OF ACIDIFICATION
As discussed in Section 1.1, the MAS Panel identified (1) the retention of deposited sulfur within
watersheds and (2) the supply of base cations from watersheds to surface waters as the most impor-
tant watershed processes affecting or mediating long-term surface water acidification (NAS, 1984).
These processes have become the focus of the DDRP. Factors other than sulfur retention and base
cation supply affect surface water acidification, but either were deemed by the Panel to be relatively
less important in long-term acidification or could not be addressed completely within the scope of the
DDRP due to time, budgetary, or logistical constraints. Several of these alternative factors were
discussed by Church and Turner (1986) and are summarized in Section 3 of this report.
1.2.1 Sulfur Retention
During the past decade, there has been an increased recognition that surface water acidification
is controlled not only by rates of hydrogen ion deposition but also by the mobility of associated anions
through the ecosystem. Galloway et al. (1983) and the 1984 NAS Panel identified controls on anion
mobility, specifically on sulfate adsorption, as one of the two dominant variables affecting the rate and
extent of surface water acidification by atmospheric deposition of mineral acids.
Three decades ago Nye and Greenland (1960) recognized the importance of anions as carriers
for cations in solution. The mobile anion paradigm they proposed [more recently applied to surface
water acidification (Johnson and Cole, 1980; Seip, 1980)] suggests that a variety of processes act
more or less independently to control the concentrations of individual anions in solution, whereas
exchange and weathering processes control the relative quantities of cations. Controls on, and
changes in, anion mobility can thus be viewed as the proximate controls on rates of cation leaching
from soils and, coupled with rates of cation resupply processes, on surface water acidification.
-------
-------
Plate 1-1. Direct/Delayed Response Project study regions and sites.
-------
-------
DDRP STUDY REGIONS
Northeast
id-Appalachian
Region
DDRP Lake Study Sites
DDRP Stream Study Sites
Southern Blue Ridge
Province
-------
-------
Within the DDRP, the primary issue with regard to anion mobility lies in forecasting temporal
changes in dissolved sulfate. Sulfur retention processes are discussed further in Section 3.3; the
status and dynamics of sulfur retention are presented in Sections 7 and 9.2 and in the integrated
watershed modelling discussions in Section 10.
1.2.2 Base Cation Supply
The NAS Panel identified rates of base cation supply from watersheds as the second dominant
factor determining the rate and ultimate degree of acidification of surface waters by acidic deposition.
Base cations are supplied principally from mineral weathering (as the original source) and by cation
exchange in soils. The exchange of cations from the soil complex to the soil solution is a rapid
process, whereas the supply of base cations from mineral weathering to the exchange complex pro-
ceeds much more slowly. The balance between these rates and the rate of cation leaching by mobile
anions is a critical factor in determining the rate of soil and surface water acidification. Mineral
weathering and cation exchange are discussed further by Church et al. (1989). Projections of rates of
cation leaching from the exchange complex are presented in Section 9.3 and are incorporated in
watershed modelling studies presented in Section 10.
1.3 GENERAL APPROACH
As H.B.N. Hynes (1975) once noted, "We must not divorce the stream from its valley in our
thoughts at any time. If we do we lose touch with reality." Although surface waters can be affected by
acidic deposition originating from emissions many miles distant, the concept of the watershed as a
unit is critical in understanding current and future aquatic effects. Indeed, for drainage lake and
reservoir systems in the Northeast, Upper Midwest, and Southern Blue Ridge Province, most ANC pro-
duction occurs as a result of biogeochemical processes within the surrounding watershed (Church et
al., 1989; Shaffer and Church, 1989; Shaffer et al., 1988). This is undoubtedly true for stream reach
systems studied in the Mid-Appalachian Region as well.
Because of the importance of watershed processes (especially those occurring in soils) in deter-
mining future aquatic effects, new data on these processes and on related soil pools and capacities
were required. Initially, we considered using existing regional soils data for the DDRP analyses.
Existing soils databases, however, were limited with respect to their application to issues of surface
water acidification. First, such data are available primarily from lowland agricultural regions, whereas
surface water acidification occurs principally in relatively undisturbed upland systems. Second, such
databases generally do not include a number of key variables relevant to soil chemical interactions
with acidic deposition.
We subsequently decided that the Project required a new regional soils database, necessitating
a major soil survey (Sections 5.1 - 5.5; also see Lee et al., 1989; Church et al., 1989). We further con-
cluded that this survey should allow the specific soils (and specific soil types) to be linked with the
NSWS databases that describe the chemistry of low ANC lakes and streams. Thus we adopted the
approach outlined in this section and illustrated in Figure 1-1.
-------
ProjecfDesign
Watershed Selection
Watershed Mapping
Soil Sampling and
Field Measurements
Soil Preparation
Chemical/Physical
Laboratory Analysis
Data Analysis
Reporting
Development of Soil *
Sampling Classes
Supporting Regional
Datasets
,.
Database Management
Figure 1-1. Steps of the Direct/Delayed Response Project (DDRP) approach. Asterisks
denote steps that received significant support from Geographic Information
Systems (GIS) activities (Campbell and Church, 1989; Campbell et al., 1989).
-------
1.3.1 Watershed Selection
We selected DDRP watersheds as a high-interest subset of lake and stream systems surveyed in
the NSWS (for details for the M-APP see Section 5.2). The watersheds are a subset of a probability
sample (i.e., the National Stream Survey - see Kaufmann et al., 1988). This allows results of DDRP
analyses to be extrapolated to a specified target population (see Section 6).
1.3.2 Soil Survey
The USDA Soil Conservation Service (SCS) prepared maps of soils, vegetation, land use, and
depth to bedrock for each DDRP watershed (see Section 5.4). Soil sample classes were defined for
each DDRP region, and soils selected from these classes were sampled and analyzed for physical and
chemical characteristics. We aggregated soils within sampling classes to develop characterizations
(e.g., class means and variances) that we used to "rebuild" or represent (e.g., by mass or area
weighting) the characteristics of study watersheds. We provide details of the sample class selection,
sampling, and soil analysis in Section 5.5.
1.3.3 Other Regional Datasets
The regional nature of the Project required estimates of precipitation, atmospheric deposition
(wet and dry), and surface water runoff (as runoff depth) developed in a standardized manner across
the eastern United States. Because study sites for the DDRP were selected as a subset of a statistical
sample, most sites had no direct information for deposition and runoff. We describe the development
of these datasets for the DDRP in Sections 5.6 and 5.7, respectively.
1.3.4 Scenarios of Atmospheric Deposition
The major question driving the DDRP concerns the response of surface water chemistry to
atmospheric deposition in the future. For the DDRP studies in the M-APP, the EPA's Office of Air and
Radiation requested that we evaluate three sulfur deposition scenarios. The first deposition scenario
for the region was that of constant deposition at current (1985) levels. A second scenario was that of
a ramp decrease beginning in five years and leveling off at 15 years at 50 percent of current. The
third scenario was an identical decrease to 15 years followed by a ramp increase for 35 years to a
level 20 percent below current (see Figure 1-2).
1.3.5 Data Analysis
We performed a variety of complementary data analyses within the Project (see Section 4.4 for
more details). The most basic of these analyses is the statistical evaluation of interrelationships
among atmospheric deposition, mapped watershed characteristics, soil chemistry, and current surface
water chemistry. We performed these extensive analyses for the NE and SBRP (Church et al., 1989).
The principal result of these analyses was to verify that the processes and relationships incorporated
in the modelling analyses reasonably represented the systems under study (see Section 8 of this
-------
Change from 1985 (X)
20 30 40
Time, yrs.
Figure 1-2. Sulfur deposition scenarios for the M-APP Region.
-------
report for a Summary discussion), thus we did not repeat such analyses for the Mid-Appalachian
Region, inasmuch as it would not have been a very productive use of our time.
Watershed retention of atmospherically deposited sulfur is an important consideration within the
Project. Current regional retention is evaluated in Section 7, and the dynamics of retention via soil
sulfate adsorption are considered in Section 9. Section 9 also contains results of the application of
single-factor models (Bloom and Grigal, 1985; Reuss and Johnson 1985, 1986) of the influence of
acidic deposition on the supply of base cations from soils to surface waters. The purpose of this type
of modelling is to evaluate the importance of cation exchange as a process mediating surface water
acidification.
Watershed models have been used in the DDRP to project integrated effects of atmospheric
sulfur deposition on surface water chemistry (Church et al., 1989). For the NE and SBRP, we applied
three models developed specifically to investigate the effects of acidic deposition on watersheds and
surface waters: (1) the Model of Acidification of Groundwater in Catchments (MAGIC) (Cosby et al.,
1985a,b; 1986a,b), (2) the Enhanced Trickle-Down (ETD) Model (Lee, 1987; Nikolaidis et al., 1988;
Schnoor et al., 1986), and (3) the Integrated Lake-Watershed Acidification Study (ILWAS) Model (Chen
et al., 1983; Gherini et al., 1985). The three models used common datasets for forcing functions (e.g.,
rainfall, runoff, atmospheric deposition) and data aggregated from the DDRP soils database for state
variables (e.g., soil physical and chemical variables). Although absolute values of concentrations of
constituents varied (in some cases, considerably) in comparisons for individual watersheds, projec-
tions of changes in ANC among the models on a regional basis (the thrust of the DDRP) were remark-
ably consistent (Church et al., 1989). Rather than repeat costly and time consuming duplicative
modelling analyses for the M-APP, we used only the MAGIC model for projections for this region. As
applied by the model developers themselves for the NE and SBRP DDRP studies, this model was by
far the easiest, quickest, and least expensive to use. We present results of our modelling analyses for
the M-APP in Section 10.
1.4 RESULTS
This section presents an overview of the results of the DDRP analyses. Church et al. (1989)
performed statistical analyses of the interrelationships among deposition, edaphic factors, and surface
water chemistry for lake/watersheds in the NE and stream/watersheds in the SBRP. Those analyses
generally supported the postulated relationships incorporated into both the single factor models (for
sulfate adsorption and cation supply) and the integrated watershed models. For example, soil depth,
soil chemical characteristics, and watershed hydrology factors all appear as important explanatory
variables in the regressions that we performed. Additionally, wetlands in northeastern watersheds
appear to have an important role in influencing sulfur dynamics (see Sections 7 and 8 of Church et al.,
1989). The watersheds we studied in the M-APP, however, contain few wetland areas (see Sections 5
and 7), and such effects are thus relatively unimportant in this region. As discussed in the Preface
and in Section 8, we did not perform the lengthy statistical analyses for the watersheds of the M-APP
Region.
-------
1.4.1 Retention of Atmospherically Deposited Sulfur
1.4.1.1 Current Retention
We computed net annual retention of atmospherically deposited sulfur for watersheds not
having apparent significant internal sources of sulfur (see Section 7). At present, watersheds in the NE
appear to be approximately at steady state, whereas median net retention is about 75 percent in the
SBRP. These observations are qualitatively consistent with theory (Galloway et al., 1983).
The M-APP Region appears to be a zone of transition, in both space and time, in terms of cur-
rent sulfur retention status. Spatially, the region is between NE systems that are, on average, near
steady state and the SBRP, in which median retention exceeds 75 percent. Current retention in
M-APP systems ranges widely, from -83 to +90 percent, with medians of 40 and 3 percent in sub-
regions 2Bn and 2Cn, respectively, (see Plate 1-2). Soils in the region have substantial sulfate adsorp-
tion capacities, and are projected to respond, in most cases, over a period of several decades to
changes in sulfur deposition (Section 9.2). Modelling results suggest that most watersheds in the
region at one time retained (and in a few cases continue to retain) the great majority of the elevated
sulfur deposition loadings, at levels comparable to those currently observed in the SBRP. The model-
ling results also suggest that continued deposition to the region is saturating soils relative to their
adsorption capacities, and that sulfate is now breaking through soils in most of the region as systems
begin to approach steady state.
There are also spatial patterns of sulfur retention within the region that might be related to differ-
ences in historic sulfur deposition and/or to current loading patterns. In the westernmost part of the
region (Subregion 2Cn of the NSS), median sulfur retention is now only three percent. High sulfur
deposition and decreased retention may have led to the low ANC and acidic stream reaches (exclud-
ing stream reaches affected by acid mine drainage) identified there by the National Stream Survey
(Kaufmann et al., 1988).
1.4.1.2 Projected Retention
As indicated above, current sulfur retention in M-APP watersheds is highly variable, ranging from
systems already at steady state to sites with retention exceeding 70 percent. The projected dynamic
response of sulfur in M-APP watersheds is similarly variable; some watersheds have already reached
steady state, yet others are not projected to come to steady state under current deposition for 150
years. The sulfate dynamics described below are based on projections of the integrated MAGIC
model, but projections from the Level II sulfate modelling (Section 9.2) are almost identical. For a
scenario of constant deposition, model projections underestimate current sulfur retention; median
simulated retention in year 0 was 26 percent (compared to 44 percent for measured retention). Reten-
tion is projected to decrease sharply during the simulation period, to a median of 12 percent in 20
years, and to only 4 percent and 2 percent by years 50 and 100, respectively (Plate 1-3). Most of the
projected decreases in retention occur during the first 50 years of simulations; the subsequent rate of
decrease in retention (increase in sulfate concentration) slows substantially because most systems are
near steady state. Median time to sulfur steady state in the M-APP, based on Level II analyses, is
10
-------
Plate 1-2. Sulfur retention and wet sulfate deposition for National Surface Water Survey
subregions in the eastern United States.
11
-------
-------
NSUIS SUBREGIONS
MEDIAN PERCENT SULFUR RETENTION
AND WET SULFATE DEPOSITION
2-25
MEDIAN PERCENT
SULFUR RETENTION
< 0
13 o - 20
Q 20 - 40
Q 40 - 60
0 60 - 80
El 80 - 100
Average Annual
Wet Sulfate t 2.75-
Deposition (g nf2 yr~')* 3.06-.
3.25
2-25
Eastern Lake Survey
Subreglon % Retention
1A
B
C
D
IE
-U
8
-9
-12
2.00
Notional Stream Survey
lied ion
Subregion % Retention
2Cn
2Sr,
38
2X
2As
3A
3
40
34
50
75
78
'Deposition for 1980 - 1984
(A. 01 sent Personal Communication)
-------
-------
considerably shorter (35 years), than in the SBRP watersheds (61 years) characterized by Church et al.
(1989).
The changes in sulfur retention represent large projected changes in sulfate concentration
during the simulation period. Levels of sulfur deposition in the M-APP are much higher than in the NE
or SBRP (Section 5.6); steady-state sulfate concentrations in the M-APP, under a scenario of constant
deposition, are also much higher (median = 215^eq/L compared to 111
-------
-------
Plate 1-3. Changes in sulfur retention up to 100 years for watersheds of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 1.4.3).
-------
-------
Mid-Appalachian
Streams
PERCENT
SULFUR RETENTION
ode! = MAGIC
3rd Quarlile 4
(1.5 x Interquartile Range) '
3rd Quartile
Mean
Median
1st Quartile
Isl Quartile -
(1.5 x Interquartile Range)
Not to exceed extreme value.
<=>
Current Deposition
50 —
0 —
-50
-100
>- GZ
50% Decrease
50 —
cc:
cc:
0 -
-50
-100
50%/20% Decrease
50
0 -
-50
'-100
O-4 IO
Q£
>— CtC
-------
-------
output exceeds weathering rates, exchange can provide a temporary source of bases to neutralize
acidity but, with time, the finite exchange reservoir will be depleted and runoff will eventually become
acidic.
Our analyses suggest that relative to the control of cation concentrations and ANC, cation
exchange is most important in systems with low surface water ANC (< 100 peq/L), and that in systems
with higher ANC, weathering resupply will maintain positive surface water ANC. In low ANC systems,
the relative contributions of weathering and exchange cannot be accurately determined.
1.4.2.2 Future Effects
Using two soil chemistry models, developed by Reuss and Johnson (1986) and by Bloom and
Grigal (1985), we assessed the potential importance of soil cation exchange in delaying surface water
acidification. These analyses represent a worst-case assessment that considers only exchange
processes in the upper two meters of soil (i.e., weathering is not considered) and projections were
generated assuming sulfate to be completely mobile. Model projections reflect two processes occur-
ring within the soil: (1) a mobile anion effect (i.e., to maintain charge balance, concentrations of both
basic and acidic cations increase or decrease with changes in sulfate) and (2) depletion of the soil
base cation exchange pool (i.e., as the proportion of base cations on the solid phase is depleted, the
proportion of bases in solution also decreases and the proportion of acidic cations in solution
increases). Results of simulations from the two cation models are generally consistent. For a
scenario of constant deposition, ANC is projected to decrease only slightly during the first 50 years of
simulations, then to drop sharply between years 50 and 100 as the exchange pool becomes depleted
(median projected decreases 0.3, 3, and 73 ^eq/L at 20, 50, and 100 years). For the same period, soil
base saturation is projected to decline steadily, from 14 to 9 percent; soil pH drops slightly during the
first 50 years of simulations, but the decline then accelerates as the exchange buffer system is
depleted. Soil solution aluminum concentrations increase in a corresponding nonlinear pattern, from
9 ^mol/L at the start of simulations to 22 jumol/L at year 100. For the scenarios of reduced sulfur
deposition, changes are much smaller. ANC is projected to increase (median increase of 6 fieq/L after
100 years) in the 50 percent decreased deposition scenario, due to a mobile anion effect. Soil pH,
base saturation, and soil solution aluminum remain almost constant under the same scenario. For the
50/20 percent decreased deposition scenario, ANC is projected to initially increase, then to decrease
by a median of 8 /*eq/L after 100 years. For the same scenario, soil pH and base saturation decrease
over the 100-year simulation period, but by much smaller amounts than if current rates of deposition
are maintained.
1.4.3 Integrated Effects of Surface Water ANC
As noted in the Preface of this report, as discussed in Section 1.3.4, and as fully described in
Section 10, MAGIC was the watershed model that we applied to study the stream watersheds in the
M-APP region. At present it is not possible to quantify the long-term (i.e., 50-100 years) accuracy of
future projections of any watershed acidification model, including the MAGIC model. Calibration and
short-term (10 years) verification tests, as well as comparisons to other field and laboratory studies
and paleoecological reconstructions of past lake chemistries, however, are reasonably encouraging
(see Sections 10 and 1.5).
14
-------
The MAGIC model was successfully calibrated for 29 of the 36 DDRP stream reaches (at the
lower nodes; see Section 10) in the M-APP. The 36 DDRP watersheds represent 5,496 stream
reaches with a combined reach length of 15,239 km. The 29 watersheds simulated by MAGIC
represent 4,298 stream reaches totalling 11,246 km of stream reach length. The modelling projections
presented here directly pertain only to the target population of watershed modelled using MAGIC. As
discussed in Section 5.2, these stream reaches are a high-interest subset of those surveyed previously
by the National Stream Survey (Kaufmann et al., 1988).
We performed modelling estimates of effects on stream chemistry for time periods up to 100
years. As discussed in Section 10, the modelling of the stream reaches, based on calibrations to
observed chemical concentrations at the lower nodes of each stream reach, results in projections of
conditions for the most buffered part (i.e., the lower end) of the stream reaches under study (see also
Kaufmann et al., 1988). Therefore, projections of changes in chemical concentrations for stream reach
lengths are conservative (i.e., underestimates).
Calibrated sulfate concentrations exceed observed concentrations, which also has an effect (see
Table 10-6). For projections of increased sulfate concentration, the change from the calibrated value
to the steady-state value will be an underestimate, leading to an underestimate of acidification (i.e.,
loss of ANC). For projections of decreased sulfate concentrations, the effect will be an overestimate of
ANC recovery.
As discussed in Section 10, the watershed modelling analyses make use of watershed soil
representations as aggregated from the DDRP Soil Survey. Because the focus of the DDRP is on
regional characteristics and responses, soils data were gathered and aggregated so as to capture the
most important central tendencies of the systems under study. As a result, the analyses presented
here underestimate extremes of individual watershed responses. Those systems that MAGIC projects
to respond to the greatest extent or most quickly to current or altered levels of sulfur deposition might,
in fact, respond even more extensively or more quickly than indicated here. This should be kept in
mind when evaluating the results of the MAGIC simulations.
The projected changes in the acid-base status of M-APP streams can be presented in a variety
of ways. Plate 1-4 illustrates the projected changes in the distributions of ANC for each of the three
deposition scenarios for up to 100 years. Plate 1-5 shows the associated projected changes in pH.
More detail on the projected changes in ANC and pH at 50 years is shown in Plates 1-6 and 1-7,
respectively. Plates 1-4 and 1-5 indicate patterns of (1) a monotonic decrease in both ANC and pH for
the constant deposition scenario, and (2) an increase in ANC and pH in the short term (20 years), with
longer term decreases for both the decreased deposition scenarios. For the scenario of 50 percent
decrease in deposition, there are slight decreases in ANC and pH over the long term, presumably due
to modelled depletion of the soil cation exchange complex. For the 50/20 percent decreased
deposition scenario, the decreases are due principally to the ramped increase of deposition back up
to the 20 percent level at year 50. Plates 1-6 and 1-7 show that, although the greatest ANC changes
are not projected for the lowest ANC group (i.e., < 25jweq/L), the largest pH changes are projected
for this group. This reflects, of course, the logarithmic relationship between these two variables.
15
-------
Plate 1-4. Changes in ANC up to 100 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see, Section 1.4.3).
16
-------
-------
id-Appalachian
Streams
ANC vs. TIME
Model = MAGIC
300
^ 200
cr
= 100
Current Deposition
0 —
-100
3rd Quartile 4
(1.5 x Interquartile Range)
3rd Quartile
Mean
Median
1st Quarlile
1st Quarliie -
(1.5 x Interquartile Range)"
" Not to exceed extreme value.
50% Decrease
C-i
300 — |
200 —
100 —
1 n n
_j
•, .
:J
fe
If
-j
_j
-i
i
cxj in
300
: 200
'••f 100
; o
-100
50%/20% Decrease
IO
eel
-------
-------
Plate 1-5. Changes in pH up to 100 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 1.4.3).
17
-------
-------
Mid-Appalachian
Streams
pH vs. TIME
Model = MAGIC
8.0
7.0
6.0
5.0
4.0
Current Deposition
O C3
3rd Quottile +
(1.5 x Interquartile Range)
3rd Quartile
Mean
Median
1st Quartile
1st Quortile -
(1.5 x Interquartile Range)
" Not to exceed extreme value.
8.0-i
7.0 -
6.0-
5.0 —
4.0-
50% Decrease
8.0 -
7.0 -
6.0-
5.0—1
4.0 -
50%/20% De-crease
-------
-------
Plate 1-6. Changes in median ANC at 50 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 1.4.3).
18
-------
-------
CHANGE IN MEDIAN ANC
Year 0 to Year 50
Model = MAGIC
M-APP Study Area
as
a>
en
Constant
Deposition
30-i
20-
10-
n
U
10-
20-
30-
§^,.,j
W.-—1
-7
v
-3
•- ,
-23
MO
NX
CD
CD
CD
IO CD
OJ CD
50%
Decrease
30n
20-
10
— PCf?
10-
20-
30-
24
i i
UO C3
50%/20%
Decrease
30n
0 A
L U
10-
A
U
10-
20-
30-
0.5
I 1
.7 -5
UO
C-J
-«*•
I
LO
04
ANC Group (ueq/L)
-------
-------
Plate 1-7. Changes in median pH at 50 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 1.4.3).
19
-------
-------
CHANGE IN MEDIAN pH
Year 0 to Year 50
Model = MAGIC
M-APP Study Area
Constant
Deposition
1.0-1 1,0
50%
Decrease
0.5H
0
C->
-0 , 5-
ex.
-0.11
-0.40
-0,29
to
V
I I
tra CD
Osl CD
0,5
0
-0,5-
0.05
0.09
to
og
cr>
o
1 I
to o
OJ O
ANC Group (ueq/L)
50%/20%
Decrease
1 ,0
0,5
0
0.5H
0.05
-0.05 -0-02
to
V
1 I
to a
Cxi O
-------
-------
Another way to present the results of the modelling is as changes in the percent or length of
stream reaches projected to have ANC < 0 ^eq/L (Figures 1-3 and 1-4) or < 50 [teq/L (Figures 1-5
and 1-6) in the future. Because MAGIC has some bias in its calibration to year 0 (see Section 10),
these values are computed by adding the projected changes in ANC to the values observed by
Kaufmann et al. (1988) for the National Stream Survey. The changes in ANC projected by MAGIC are
unbiased with respect to year-0 ANC (J. Cosby, pers. comm.). A one-sided hypothesis test of sig-
nificance for the changes in numbers of systems (i.e., different from zero; see Section 6) indicates
changes (p < .05) at years 20, 50, and 100 for stream reaches with ANC < 0 ^eq/L for the current
deposition scenario (see Section 10). Changes at 50 and 100 years for the projections of systems
with ANC < 50 fteq/L are also evident (p = .01). For the 50/20 percent decreased deposition
scenario, changes (increases) in the numbers of stream reaches with ANC < 0 or ANC < 50 fieq/L are
indicated (.07 .16) (see Table 10-9).
Considered together, the results shown in Plates 1-4 through 1-7 and Figures 1-3 through
1-6 indicate a modelled projection of marked decreases in the ANC and pH of M-APP stream reaches
(as represented by chemical conditions at the lower reach nodes) under continued levels of current
deposition. These projected changes apply to chemical conditions at lower reach nodes and only to
the target population represented by the systems modelled by MAGIC. These projected changes in
ANC translate to marked increases in (1) the number of acidic stream reaches (ANC < O^eq/L), and
(2) the number of stream reaches with ANC < 50 ^eq/L Stream reaches with ANC < 50 fieq/L are
probably highly susceptible to episodic acidification to acidic conditions during periods of heavy
snowmelt or rainfall (Eshleman, 1988; Wigington et al., 1990).
Projections for the 50 percent decreased deposition scenario indicate somewhat improved
(relative to toxic conditions for fish) chemical conditions at 50 years (especially pH for low ANC
systems), with perhaps a slight decrease in water quality after that time. Projections for the 50/20
percent decreased deposition scenario indicate chemical conditions at 50 years very close to those
observed by Kaufmann et al. (1988) in the National Stream Survey. Decreases in water quality are
projected at 100 years under this deposition scenario.
1.5 SUMMARY DISCUSSION
Our analyses of (1) current sulfur mobility, (2) projected sulfur mobility, (3) cation exchange
buffering, and (4) integrated watershed modelling present an internally consistent picture of the
potential long-term effects of sulfur deposition on the chemistry of our target population of stream
reaches in the M-APP region. Although soils of our study watersheds have greater ability to retain
future atmospheric inputs of sulfate than do those of the NE, this ability is quite limited relative to soils
of the SBRP (Church et al., 1989). Soils of the M-APP Region are projected to reach sulfur input-
output steady state faster than those of the SBRP (median time to steady state is 35 years for the
M-APP, 61 years for the SBRP). Calculated steady-state sulfate concentrations in the M-APP Region
(214^teq/L) are substantially greater than for the NE (111 /ueq/L) or SBRP (120|*eq/L). Leaching of
base cations from these soils also is projected to play an important role in future acidification in the
20
-------
30
o
O)
cc
•4—•
c
0)
o
10-
Current Deposition
in
CO
Y///7A
0
20
50
100
iJU"
CO 20-
C
Q)
O
o 10-
D.
•
o-
50% Decreased Deposition
CO
Y/////, 000
0 20 50 100
JU ~
o
D)
CO 20-
c
o
o
% 10-
o-
50%/20% Decreased Deposition
co "*•
Y/////A ° y/////,
0 20 50
CO
H
100
Time (yr)
Figure 1-3. Percentage of stream reaches of the DDRP M-APP modelling target population
(see Section 1.4.3) projected by the MAGIC model to have ANC < 0 ueq/L at
the lower reach nodes. Estimates at year 0 are from the National Stream
Survey (Kaufmann et al., 1988) for the same target population. Estimates at
20, 50, and 100 years are computed by adding the change in ANC projected
by the MAGIC model to ANC observed by the National Stream Survey.
21
-------
s_x
JC
o
o
(0
(1)
cc
3000
2000-
1000-
3000
£j 20001
c
o
1000 -
Current Deposition to
00
CO
co T"
o>
CM
CO
CM
w//,
20
50
100
50% Decreased Deposition
CM
0
20
50
100
3000-
2000-
1000-
n _
50%/20% Decreased Deposition
<£>
O
in T-
CN| "i1 'W/'/1/1
** ^77777 f
-------
100
a 7£
rc
I 5 25-
Current Deposition
CO
CM
20
m
50
CO
in
100
o
o
CO
*-•
£
O
o
k.
-------
c
CD
o
CO
O
DC
10000
7500-
5000-
2500-
0
Current Deposition
in
o
10
CM
CO
00
in
CM
CO
(O
20
50
100
D)
C
O
O
03
0>
DC
10000
7500
5000
50% Decreased Deposition
m
o
in
CM
CM
CM
CM
CM
CM
CM
CO
O)
•ef
CM
20
50
100
10000
D)
C
O
£
o
(0
0)
DC
7500-
5000-
2500-
0
50%/20% Decreased Deposition
o>
(O
o
nn
CM
CM
CM
CM
CM
0 20 50 100
Time (yr)
Figure 1-6. Length (km) of stream reaches of the DDRP M-APP modelling target popula-
tion (see Section 1.4.3) projected by the MAGIC model to have ANC < 50
ueq/L at the lower reach nodes. Estimates °at year 0 are from the National
Stream Survey (Kaufmann et al., 1988) for the same target population. Esti-
mates at 20, 50, and 100 years are computed by adding the change in ANC
projected by the MAGIC model to ANC observed by the National Stream Survey.
24
-------
region, especially after 50 years. Integrated watershed modelling projects that many more systems in
the M-APP than in either the NE or SBRP may become chronically acidic or may become susceptible
to episodic acidification in the next 50 to 100 years under current levels of sulfur deposition. Model
projections indicate that a decrease in atmospheric sulfur deposition of about 50 percent would
prevent such changes over the 50 to 100 year period simulated.
An ever increasing amount of laboratory experimental data (Dahlgren et al., 1990), observational
field data from lakes (Dillon et al., 1986; Hutchinson and Havas, 1986; Keller and Pitblado, 1986) and
streams (Swank and Waide, 1988; Ryan et al., 1989; Morgan, 1990), and data from manipulated
catchments (Wright et al., 1988) support both the theory (Galloway et al., 1983) and the qualitative
(and to an extent, quantitative) future changes projected by our studies.
1.6 REFERENCES
Altshuller, A.P., and R.A. Linthurst. 1984. The Acidic Deposition Phenomenon and Its Effects: Critical
Assessment Review Papers. EPA/600/8-83/016bf. U.S. Environmental Protection Agency,
Washington, DC.
Asbury C.E., F.A. Vertucci, M.D. Mattson, and G.E. Likens. 1989. Acidification of Adirondack lakes.
Environ. Sci. Technol. 23:362-365.
Baker, LA., P.R. Herlihy, P.R. Kaufmann, and J.M. Eilers. 1991. Acidic lakes and streams in the United
States: The role of acidic deposition. Science. 252:1151-1154.
Bloom, P.R., and D.F. Grigal. 1985. Modeling soil response to acidic deposition in nonsulfate
adsorbing soils. J. Environ. Qual. 14:489-495.
Campbell, W.G., and M.R. Church. 1989. EPA uses GIS to study lake and stream acidification.
Federal Digital Cartography Newsletter 9:1-2.
Campbell, W.G., M.R. Church, G.D. Bishop, D.C. Mortenson, and S.M. Pierson. 1989. The role for a
geographic information system in a large environmental project. Internal. J. GIS. 3:349-362.
Chen, C.W., S.A. Gherini, J.D. Dean, R.J.M. Hudson, and R.A. Goldstein. 1983. Modeling of
Precipitation Series, Volume 9. Ann Arbor Sciences, Butterworth Publishers, Boston, MA. 175 pp.
Church, M.R. 1989. Predicting the future longrterm effects of acidic deposition on surface water
chemistry: The Direct/Delayed Response Project. Eos, Transactions, American Geophysical
Union 801-802, 812-813.
Church, M.R., and R.S. Turner. 1986. Factors Affecting the Long-term Response of Surface Waters to
Acidic Deposition: State of the Science. EPA/600/3-86/025. NTIS PB 86 178 188-AS. U.S.
Environmental Protection Agency, Corvallis, OR. 274 pp.
25
-------
Church, M.R., K.W. Thornton, P.W. Shaffer, D.L Stevens, B.P. Rochelie, G.R. Holdren, M.G. Johnson,
J.J. Lee, R.S. Turner, D.L. Cassell, D.A. Lammers, W.G. Campbell, C.I. Liff, C.C. Brandt, LH.
Liegel, G.D. Bishop, D.C. Mortenson, S.M. Pierson and D.D. Schmoyer. 1989. Direct/Delayed
Response Project: Future effects of long-term sulfur deposition on surface water chemistry in
the Northeast and Southern Blue Ridge Province. EPA/600/3-89/061 a-d, Washington, DC, 887
pp.
Cosby, B.J., G.M. Hornberger, J.N. Galloway, and R.F. Wright. 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.
Cosby, B.J., G.M. Hornberger, J.N. Galloway, and R.F. Wright. 1985b. Time scales of catchment
acidification: A quantitative model for estimating freshwater acidification. Environ. Sci. Technol.
19:1144-1149.
Cosby, B.J., G.M. Hornberger, E.B. Rastetter, J.N. Galloway, and R.F. Wright. 1986a. Estimating
catchment water quality response to acid deposition using mathematical models of soil ion
exchange processes. Geoderma 38:77-95.
Cosby, B.J., G.M. Hornberger, R.F. Wright, and J.N. Galloway. 1986b. Modeling the effects of acid
deposition: control of long-term sulfate dynamics by soil sulfate adsorption. Water Resour. Res.
22:1283-1292.
Gumming, B.F., J.P. Smol, J.C. Kingston, D.F. Charles, H.J.B. Birks, K.E. Cambun, S.S. Dixit, A.J.
Uutala, and A.R. Selle. How much acidification has occurred in Adirondack region lakes (New
York, USA) since pre-industrial times? Can. J. Fish. Aquat. Sci. (In press).
Dahlgren, R.A., D.C. McAvoy, and C.T. Driscoll. 1990. Acidification and recovery of a spodosol Bs
horizon from acidic deposition. Environ. Sci. Technol. 24:531-537.
Dillon, P.J., R.A. Reid, and R. Girard. 1986. Changes in the chemistry of lakes near Sudbury, Ontario
following reductions of SO2 emissions. Water Air Soil PoIIut. 31:59-65.
Driscoll, C.T., and R.M. Newton. 1985. Chemical characteristics of Adirondack lakes. Environ. Sci.
Technol. 19:1018-1024.
Driscoll, C.T., C.P. Yatsko, and F.J. Unangst. 1987. Longitudinal and temporal trends in the water
chemistry of the North Branch of the Moose River. Biogeochemistry 3:37-61.
Driscoll, C.T., N.M. Johnson, G.E. Likens, and M.C. Feller. 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.
26
-------
Driscoll, C.T., R.D. Fuller, and W.D. Schecher. 1989. The role of organic acids in the acidification of
surface waters in the eastern U.S. Water Air Soil Pollut. 43:21-40.
Eshleman, K.N. 1988. Predicting regional episodic acidification of surface waters using empirical
models. Water Resour. Res. 24:1118-1126.
Galloway, J.N., 8.A. Norton, and M.R. Church. 1983. Freshwater acidification from atmospheric
deposition of sulfuric acid: a conceptual model. Environ. Sci. Technol. 17:541-545.
Galloway, J.N., G.R. Hendrey, C.L Schofield, N.E. Peters, and A.H. Johannes. 1987. Processes and
causes of lake acidification during spring snowmelt in the west-central Adirondack Mountains,
New York. Can. J. Fish. Aquat. Sci. 44:1595-1602.
Gherini, S.A., L Mok, R.J. Hudson, G.F. Davis, C.W. Chen, and R.A. Goldstein. 1985. The ILWAS
model: Formulation and application. Water Air Soil Pollut. 26:425-459.
Hedin, L.O., G.E. Likens, K.M. Postek, and C.T. Driscoll. 1990. A field experiment to test whether
organic acids buffer acid deposition. Nature. 345:798-800.
Henriksen, A., arid D.F. Brakke. 1988. Increasing contributions of nitrogen to the acidity of surface
waters in Norway. Water Air Soil Pollut. 42:183-201.
Hutchinson, T.C., and M. Havas. 1986. Recovery of previously acidified lakes near Coniston, Canada
following reductions in atmospheric sulphur and metal emissions. Water Air Soil Pollut. 28:319-
333.
Hynes, H.B.N. 1975. Edgardo Baldi memorial lecture: The stream and its valley. Verh. Internat. Verein.
Limnol. 19:1.
Johnson, D.W., and D.W. Cole. 1980. Anion mobility in soils: Relevance to nutrient transport from
forest ecosystems. Environ. Internat. 3:79-90.
Kaufmann, P.R., A.T. Herlihy, J.W. Elwood, M.E. Mitch, W.S. Overton, M.J. Sale, J.J. Messer, K.A.
Cougar, D.V. Peck, K.H. Reckhow, A.J. Kinney, S.J. Christie, D.D. Brown, C.A. Hagley, and H.I.
Jager. 1988. Chemical Characteristics of Streams in the Mid-Atlantic and Southeastern United
States, Volume I: Population Descriptions and Physico-Chemical Relationships.
EPA/600/3-88/021 a. U.S. Environmental Protection Agency, Washington, DC. 397 pp.
Keller, W., and J.R. Pitblado. 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.
Krug, E.G. 1989. Assessment of the theory and hypotheses of the acidification of watersheds. Illinois
State Water Survey Division, SWS contract Report 457. 252 pp.
27
-------
Krug, E.G., and C.R. Frink. 1983. Acid rain on acid soil: A new perspective. Science 221:520-525.
Krug E.G., and P.J. Isaacson. 1984. Comparison of water and dilute acid treatment on organic and
inorganic chemistry of leachate from organic-rich horizons of an acid forest soil. Soil Sci.
137:370-378.
Krug, E.G., P.J. Isaacson, and C.R. Frink. 1985. Appraisal of some current hypotheses describing
acidification of watersheds. J. Air Pollut. Control Assoc. 35:109-114.
Landers, D.H., J.M. Eilers, D.F. Brakke, W.S. Overton, P.E. Kellar, M.E. Silverstein, R.D. Schonbrod,
R.E. Crowe, R.A. Linthurst, J.M. Omernik, S.A. league, and E.P Meier. 1987. Characteristics of
Lakes in the Western United States, Volume I: Population Descriptions and Physico-Chemical
Relationships. EPA/600/3-86/054a. U.S. Environmental Protection Agency, Washington, DC.
176pp.
Lee, S. 1987. Uncertainty Analysis for Long-term Acidification of Lakes in Northeastern USA. Ph.D.
Thesis. University of Iowa, Iowa City.
Lee, J.J., D.A. Lammers, M.G. Johnson, M.R. Church, D.L Stevens, D.S. Coffey, R.S. Turner, LJ.
Blume, L.H. Liegel, and G.R. Holdren. 1989. Watershed surveys to support an assessment of the
regional effect of acidic deposition on surface water chemistry. Environ. Mgt. 13:95-108.
Linthurst, R.A., D.H. Landers, J.M. Eilers, D.F. Brakke, W.S. Overton, E.P. Meier, and R.E. Crowe.
1986. Characteristics of Lakes in the Eastern United States, Volume I: Population Descriptions
and Physico-Chemical Relationships. EPA/600/4-86/007a. U.S. Environmental Protection Agency,
Washington, DC. 136 pp.
McColl, J.G., and A.A. Pohlman. 1986. Soluble organic acids and their chelating influence on Al and
other metal dissolution from forest soils. Water Air Soil Pollut. 31:917-927.
Messer, J.J., C.W. Ariss, J.R. Baker, S.K. Drouse, K.N. Eshleman, P.R. Kaufmann, R.A. Linthurst, J.M.
Omernik, W.S. Overton, M.J. Sale, R.D. Schonbrod, S.M. Stambaugh, and J.R. Tuschall Jr. 1986.
National Stream Survey Phase I, Pilot Survey. EPA/600/4-86/026. U.S. Environmental Protection
Agency, Washington, DC. 179 pp.
Mohnen, V.A. 1988. The challenge of acid rain. Scientific American 259:30-38.
Morgan, M.D. 1990. Streams in the New Jersey Pinelands directly reflect changes in atmospheric
deposition chemistry. J. Environ. Qual. 19:296-302.
NAS. 1984. Acid Deposition: Processes of lake acidification. Summary of Discussion. National
Research Council Commission on Physical Sciences, Mathematics, and Resources.
Environmental Studies Board, Panel on Processes of Lake Acidification. National Academy
Press, Washington, DC. 11 pp.
28
-------
NAS. 1986. Acid Deposition: Long-Term Trends. National Research Council Commission on Physical
Sciences, Mathematics, and Resources. Environmental Studies Board. National Academy
Press, Washington, DC. 506 pp.
Neary, B.P., and P.J. Dillon. 1988. Effects of sulphur deposition on lake-water chemistry in Ontario,
Canada. Nature 333:340-343.
Nikolaidis, N.P., H. Rajaram, J.L Schnoor, and K.P. Georgakakos. 1988. A generalized soft water
acidification model. Water Resour. Res. 24:1983-1996.
Nye, P.H., and D J. Greenland. 1960. The Soil Under Shifting Cultivation. Commonwealth Bureau of
Soils Tech. Comm. No. 51. Commonwealth Agricultural Bureaux, Farnham Royal, Bucks.
Reuss, J.O., and D.W. Johnson. 1985. Effect of soil processes on the acidification of water by acid
deposition. J. Environ. Qual. 14:26-31.
Reuss, J.O., and D.W. Johnson. 1986. Acid Deposition and the Acidification of Soils and Waters.
Ecological Studies Volume 59. Springer-Verlag, Inc., New York, NY.
Ryan, P.P., G.M. Homberger, B.J. Cosby, J.N Galloway, J.R. Webb, and E.B. Rastetter. 1989.
Changes in the chemical composition of stream water in two catchments in the Shenandoah
National Park, Virginia, in response to atmospheric deposition of sulfur. Water Resour. Res.
25:2091-2099.
Schindler, D.W. 1988. The effects of acid rain on freshwater ecosystems. Science 239:149-157.
Schnoor, J.L, N.P. Nikolaidis, and G.E. Glass. 1986. Lake resources at risk to acidic deposition in the
Upper Midwest. J. Water Pollut. Control Fed. 58:139-148.
Seip, H.M. 1980. Acidification of freshwaters - sources and mechanisms, pp. 358-366. In: D. Drablos
and A. Tollan, eds. Ecological Impact of Acid Precipitation, Proceedings of an International
Conference, Sandefjord. March 11-14. SNSF Project, Oslo-As, Norway.
Shaffer, P.W., and M.R. Church. 1989. Terrestrial and in-lake contributions to the alkalinity budgets of
drainage lakes: An assessment of regional differences. Can. J. Fish. Aquat. Sci. 46:509-515.
Shaffer, P.W., R.P. Hooper, K.N. Eshleman, and M.R. Church. 1988. Watershed vs. in-lake alkalinity
generation: A comparison of rates using input-output studies. Water Air Soil Pollut. 39:263-273.
Sullivan, T.J., J.M. Eilers, M.R. Church, D.J. Blick, K.N. Eshleman, D.H. Landers, and M.S. DeHaan.
1988. Atmospheric wet sulphate deposition and lakewater chemistry. Nature 331:607-609.
29
-------
Swank, W.T., and J.B. Waide. 1988. Characterization of baseline precipitation and stream chemistry
and nutrient budgets for control watersheds, pp. 57-79. In: W.T. Swank and D.A. Crossley, Jr.,
eds. Ecological Studies, Volume 66: Forest Hydrology and Ecology at Coweeta. Springer-Verlag,
New York, NY.
Turner, R.S., R.B. Cook, H. Van Miegroet, D.W. Johnson, J.W. Elwood, O.P. Bricker, S.E. Lindberg,
and G.M. Hornberger. 1990. Watershed and Lake Processes Affecting Chronic Surface Water
Acid-Base Chemistry, NAPAP SOS/T Report 10, In: Acidic Deposition: State of Science and
Technology. National Acid Precipitation Assessment Program, Volume II, Washington, DC.
Wigington, P.J., Jr., T.D. Davies, M. Tranter, and K.N. Eshleman. 1990. Episodic Acidification of
Surface Waters Due to Acidic Deposition. NAPAP SOS/T Report 12. In: Acidic Deposition: State
of Science and Technology. National Acid Precipitation Assessment Program, Volume II,
Washington, DC.
Wright, R.F., E. Lotse, and A. Semb. 1988. Reversibility of acidification shown by whole-catchment
experiments. Nature 334:670-675.
30
-------
-------
SECTION 2
INTRODUCTION TO THE DIRECT/DELAYED RESPONSE PROJECT
2.1 PROJECT BACKGROUND
Much scientific interest and public debate surround the subject of the effects of acidic deposition
on freshwater ecosystems (e.g., Schindler, 1988; Mohnen, 1988). The U.S. Environmental Protection
Agency (EPA) recently completed a comprehensive chemical survey, the National Surface Water
Survey (NSWS), of lakes and streams in the United States considered to be most vulnerable to acidic
deposition (i.e., those with the lowest acid neutralizing capacity, or ANC) (Linthurst et al., 1986;
Landers et al., 1987; Kaufmann et al., 1988). Analysis of these and other lake and stream chemistry
data, together with data on temporal and spatial patterns of atmospheric deposition, indicates that
long-term deposition of sulfur-containing compounds originating from the combustion of fossil fuels
has acidified (i.e., decreased the ANC of) some surface waters in eastern North America (Altshuller
and Linthurst, 1984; NAS, 1986; Sullivan et al., 1988; Neary and Dillon, 1988; Asbury et al., 1989;
Baker et al., 1991).
The transport of mobile anions (primarily sulfate) through watershed soils and closely associated
cation leaching are the most widely accepted mechanisms of this acidification process (Seip, 1980;
Galloway et al., 1983; Driscoll and Newton, 1985; Church and Turner, 1986). Krug (see Krug and
Frink, 1983; Krug et al., 1985; Krug, 1989) has hypothesized that acidic deposition may have shifted
the nature of some very low ANC or naturally acidic surface waters from organic acid "dominance" to
mineral acid dominance. A number of studies have examined or have shed light upon this question
from the perspective of soil processes (e.g., Krug and Isaacson, 1984; McColl and Pohlman, 1986),
observations of surface water chemistry (Driscoll et al., 1988; Driscoll et al., 1989a), or manipulation
experiments (Hedin et al., 1990), but have not confirmed any widespread importance of the phenom-
enon. Examination of historical changes in surface water chemistry of Adirondack lakes (as recon-
structed from patterns of changes in diatom and chrysophyte populations), in fact, show no consistent
decreases in DOC (Gumming et al., In press) and argue against the hypothesis. In any event, the net
effect of atmospheric deposition of sulfuric acid on surface water chemistry is a shift toward aquatic
systems more dominated by mineral acidity and more likely to contain inorganic forms of aluminum,
which are toxic to aquatic organisms.
Given that acidification of some surface waters has occurred, critical scientific and policy
questions focus on whether acidification is continuing in the regions noted, whether it is just beginning
in other regions, how extensive the effects may become, and over what time scales the effects may
occur. The EPA is examining these questions through the activities of the Direct/Delayed Response
Project (DDRP) (Church and Turner, 1986; Church, 1989). The Project began in 1984 at the specific
request of the EPA Administrator, following a meeting of the Panel on Processes of Lake Acidification
of the National Academy of Sciences (NAS). Principal among the conclusions of the Panel was that
atmospheric deposition of sulfur-containing compounds is the major source of long-term surface water
acidification in eastern North America (NAS, 1984). The Panel also debated at length the dynamic
aspects of the acidification process. The DDRP was designed to focus on the topic of acidification
31
-------
dynamics and draws its name from consideration of whether acidification is immediate (immediately
proportional to levels of deposition, i.e., "direct"), or whether it lags in time (i.e., is "delayed") because
of edaphic characteristics. Church and Turner (1986) presented a discussion of the processes of
long-term surface water acidification and methods for its investigation at the beginning of the Project.
Church et al. (1989) presented a relatively brief and more current discussion of processes especially
relevant to this Project. Major points of that discussion are repeated in Section 3 of this report.
Although more recent research has indicated that deposition of nitrogen-containing compounds
may be important to both episodic (Galloway et al., 1987; Driscoll et al., 1987a; Wigington et al., 1990)
and long-term (Henriksen and Brakke, 1988) acidification of surface waters, the DDRP does not
address these effects. Such effects are the focus of developing or ongoing research within other EPA
research programs.
2.2 PRIMARY OBJECTIVES
The DDRP has four technical objectives related to atmospheric/terrestrial/aquatic interactions:
• to describe the regional variability of soil and watershed characteristics,
• to determine which soil and watershed characteristics are most strongly related to surface
water chemistry,
« to estimate the relative importance of key watershed processes in moderating regional effects
of acidic deposition, and
« to classify a sample of watersheds with regard to their response to inputs of acidic deposition
and to extrapolate the results from this sample of watersheds to the study regions.
The fourth objective is the critical "bottom line" of the Project.
The relationship of the DDRP to other projects within the Aquatic Effects Research Program
(AERP) of the National Acid Precipitation Assessment Program (NAPAP) is shown in Figure 2-1. The
DDRP was never intended to serve as a "research" project on the exact mechanisms and processes of
surface water acidification. Rather, the principal mandate of the Project was to make regional pro-
jections of the future effects of sulfur deposition on long-term surface water chemistry (principally
ANC), based upon the best available data and most widely accepted hypotheses of the acidification
process. Further watershed modelling activities within the NAPAP Integrated Assessment (see Figure
2-1) will investigate a variety of sulfur deposition scenarios and potential effects on biologically relevant
surface water chemistry (e.g., pH, calcium, and inorganic monomeric aluminum).
2.3 STUDY REGIONS
The Project focuses on three regions of the eastern United States where low ANC surface waters
are located and where levels of atmospheric deposition are greatest (relative to other U.S. regions):
(1) the Northeast (NE), (2) upland areas of the Mid-Atlantic (referred to here as the Mid-Appalachian
Region [M-APP]), and (3) the mountainous section of the Southeast called the Southern Blue Ridge
32
-------
Eastern Lake Survey
• Survey of lake index
chemistry
Watershed Manipulation Project
• Watershed process research
• Watershed manipulation by acid
addition
• DDRP model testing
T
National Stream Survey
• Survey of stream reach
index chemistry
Episodic Response Project
• Evaluation of episodic
acidification of streams
Direct/Delayed Response Project
° Projections of future effects of
long-term sulfur deposition on
surface water chemistry
- Northeastern lakes and
Southern Blue Ridge Province
streams
- Mid-Appalachian streams
State of Science
• Comprehensive analysis of
evidence for aquatic effects
Integrated Assessment
• Synthesis of aquatic effects
state of science
• Comparative evaluation of
aquatic effects for various
emissions control scenarios
Critical Loads Project
• Evaluation of critical loads
of sulfur deposition on selected
aquatic resources
Figure 2-1. Activities of the Aquatic Effects Research Program within the National Acid
Precipitation Assessment Program.
33
-------
Province (SBRP) (Plate 2-1). Initiation of the Project depended on the availability of the regional
surface water chemistry data of the National Surface Water Survey (NSWS). Thus, the Project focused
its initial work on the lake resources of the NE (Linthurst et al., 1986) and the stream resources of the
SBRP (Messer et al., 1986). Church et al. (1989) presented the DDRP results for those two regions.
Recent stream survey work by Kaufmann et al. (1988) has allowed investigation of possible future
effects of acidic deposition on stream chemistry in the Mid-Appalachian (M-APP) region. This report
presents the results of our work in that region.
2.4 TIME FRAMES OF CONCERN
The DDRP focuses on the potential effects of acidic deposition on surface water ANC as
evaluated at key annual "index" periods. These index periods follow the sampling schemes of the
NSWS (i.e., fall period of complete mixing for lakes and spring baseflow for streams; see Section 5.3).
"Episodic" acidification (e.g., due to snowmelt or intense rainstorms) is not considered within the
DDRP, but is the primary consideration of a companion project within the AERP, the Episodic
Response Project (Figure 2-1).
The primary time horizon for DDRP analyses is 50 years. This horizon relates to the projected
lifetimes of existing power plants and the potential implementation of additional emissions controls
relative to those lifetimes. Where possible and reasonable, some time-dependent analyses are
extended beyond this 50-year horizon to better evaluate process rates and changes and potential
future effects.
2.5 PROJECT PARTICIPANTS
The DDRP was designed and implemented at EPA's Environmental Research Laboratory -
Corvallis (ERL-C) and is a very large effort involving many participants. The Project involves two other
EPA laboratories, the Atmospheric Research and Exposure Assessment Laboratory - Research
Triangle Park (AREAL-RTP) and the Environmental Monitoring and Systems Laboratory - Las Vegas
(EMSL-LV). The DDRP is assisted by three other federal agencies: the U.S. Department of Agriculture
(including the Forest Service and the National Office, 2 National Technical Centers, and 12 state
offices of the Soil Conservation Service), the U.S. Geological Survey, and the National Oceanic and
Atmospheric Administration. Two national laboratories [Oak Ridge National Laboratory (ORNL) and
Battelle - Pacific Northwest Laboratories (PNL)], five state and private universities, and four consulting
firms also have participated in this Project. In all, over 200 field, laboratory, database management,
scientific, and management personnel have contributed to this effort.
2.6 REPORTING
This report documents and discusses the data analyses performed for the M-APP Region. It
does not contain a complete list of all data used or all results produced in the analyses. The complete
list and documentation will be available at a later date. Section 5 of this report, however, does contain
appropriate summary and example data.
34
-------
Plate 2-1. Direct/Delayed Response Project study regions and sites.
35
-------
-------
DDRP STUDY REGIONS
Northeast
Mid-Appalachian
Region
DDRP Lake Study Sites
DDRP Stream Study Sites
Southern Blue Ridge
Province
-------
-------
During the course of the Project, many of its activities have been documented, externally peer
reviewed and approved, and published as EPA reports. Any reference used in this report that has an
EPA publication number is a final, externally peer-reviewed product of this (or another) EPA project.
Such documents usually contain more complete descriptions and details of the work undertaken than
can be presented in this report. Copies of these cited EPA published reports are available upon
request from the Project Technical Director, M. Robbins Church, at ERL-C.
Project participants have published descriptions of activities and results of the DDRP in the
peer-reviewed literature. Published papers and manuscripts in press are cited throughout the report
and, like the published EPA reports, can be obtained by request from the Technical Director. As of
this writing, additional peer-reviewed publications that document the activities and results of the DDRP
are in preparation or are planned. Other preliminary results and discussions of the Project have been
presented at meetings and workshops of the American Geophysical Union (fall, 1987); American
Geophysical Union Chapman Conference on Hydrogeochemical Responses of Forested Catchments
(September 1989); Association of American Geographers (November 1987); Biometric Society (July
1986); First USA/USSR Joint Conference on Environmental Hydrology and Hydrogeology (June 1990);
International Society of Ecological Modelling (August 1987); North American Lake Management
Society (November 1986); and Soil Science Society of America (December 1987, 1988, and October
1989).
36
-------
-------
SECTION 3
PROCESSES OF ACIDIFICATION
3.1 INTRODUCTION
As discussed in Section 2.1, EPA designed and implemented the Direct/Delayed Response
Project to assess the hypotheses and conclusions of the NAS Panel on Processes of Lake Acidification
(NAS, 1984). The NAS Panel evaluated a broad range of processes potentially affecting the rate
and/or extent of surface water acidification and concluded that the most important processes were (1)
retention of deposited sulfur within watersheds (principally by adsorption) and (2) the supply of base
cations (by cation exchange and mineral weathering) from watersheds to surface waters. The panel
also considered several other processes and watershed factors, but concluded that either (1) they
were of lesser importance in controlling surface water chemistry in most systems susceptible to
acidification, and/or (2) there were insufficient data to evaluate the roles of those processes.
The design of the DDRP was intended to assess the role of time-variant processes in surface
water acidification, and more specifically to evaluate the occurrence and time frame of "delayed
response" acidification. At the start of the DDRP, there was a thorough review of watershed chemistry
processes and models (Church and Turner, 1986). A subsequent review considered recent process-
related findings vis-a-vis processes incorporated in chemistry models used in the DDRP, and was
included with DDRP analyses for the NE and SBRP (Church et al., 1989; Section 3). As part of the
final NAPAP assessment, Turner et al. (1990) have recently completed a comprehensive review and
synthesis of watershed acidification processes. This section briefly summarizes the previous process
reviews in the context of models and analyses used in the DDRP; we refer readers to the previously
cited documents for a comprehensive discussion of processes affecting surface water acidification.
3.2 THE MOBILE ANION PARADIGM -- A CONCEPTUAL MODEL OF ACIDIFICATION
During the past decade, there has been an increased recognition that surface water acidification
is controlled not only by rates of hydrogen deposition, but also by the mobility of associated anions
through ecosystems. A conceptual model of watershed acidification (Galloway et al., 1983) and the
1984 NAS panel identified both (1) controls on anion mobility, especially by sulfate adsorption, and (2)
rates of base cation supply as important variables affecting the rate and extent of acidification.
Three decades ago, Nye and Greenland (1960) recognized the importance of anions as
"carriers" for cations in solution. The "mobile anion" paradigm that they proposed, more recently
applied to surface water acidification (Johnson and Cole, 1980; Seip, 1980), suggests that a variety of
processes (e.g., adsorption of sulfate or phosphate, biologically mediated retention of nitrate, pH- and
pCO2- dependent dissociation of carbonic acid) act more or less independently to control the
concentrations of individual anions in solution, whereas weathering and cation exchange processes
control the relative quantities of cations. Controls on, and changes in, anion mobility can thus be
regarded as the proximate controls on rates of cation leaching from soils and, coupled with rates of
cation resupply processes, on surface water acidification. Within the DDRP, the primary issue with
regard to anion mobility lies in forecasting temporal changes in dissolved sulfate. Processes
37
-------
regulating mobility of sulfate and other anions are briefly discussed in Section 3.3, and the duration of
delays in acidification due to sulfate adsorption are evaluated in Section 9.2.
Cation supply from watersheds is influenced by two principal processes, mineral weathering and
base cation exchange. If cation resupply to a watershed, by deposition and mineral weathering,
exceeds the rate of strong acid anion (i.e., SO4 , NO3 , Cl) leaching, there will be net generation of
alkalinity and surface waters will not become acidic. Conversely, if the mobile anion flux exceeds the
rate of base cation resupply, exchangeable base cations can temporarily balance mobile anion
leaching demands, but the exchange pool will gradually be depleted and runoff will become
increasingly acidic as the fraction of hydrogen and aluminum increases in the leachate. The time
frame for which cation exchange can delay acidification depends on the rate of mobile anion leaching,
the magnitude of the exchange pool, and the selectivity of the exchange reactions, and can last from a
few years to many decades (Church et al., 1989, Section 9). Weathering and exchange processes are
briefly summarized in Section 3.4, and the potential for delays in acidification by exchange is evaluated
in Section 9.
3.3 ANION MOBILITY
The focus on sulfate in the NAS panel discussions and in the DDRP results from several factors.
Critical among these is the fact that sulfate (including sulfur dioxide) is the dominant anion in acidic
deposition in the eastern United States and is also the dominant anion in many low ANC surface
waters affected by acidic deposition. In addition, sulfate dynamics are time variant; sulfate can be
retained in soils by adsorption and other processes, reducing the rate of sulfate leaching and delaying
the effects of acidic deposition on surface waters. Because the duration of such delays can vary from
a few years to well over a century (Church et al., 1989), the ability to understand and predict the
temporal dynamics of sulfate is an essential objective of the DDRP. The principal sulfate retention
process in most soils on watersheds in the DDRP target population appears to be adsorption, but
alternate processes have also been considered and quantified to the extent possible.
3.3.1 Sulfate
Adsorption has long been recognized as an important process affecting sulfate mobility in soils
and the availability of sulfur to plants. Almost three decades ago, Chao and coworkers (Chao et al.,
1962a,b; 1964a,b) conducted pioneering work that identified (1) adsorption as an important retention
mechanism, (2) soil attributes affecting adsorption capacity, and (3) nonlinear isotherms as an effective
means of describing sulfate partitioning between solid and solution phases. Sulfate adsorption on
soils can occur either by nonspecific adsorption (an electrostatic bonding) or by specific adsorption, a
mechanism involving ligand exchange (with OH" or OH2) and ionic bonding, and which can involve
exchange of one or two surface ligands (Hingston et al., 1967; 1972; Rajan, 1978; 1979; Parfitt and
Smart, 1978).
Adsorption of sulfate by soils is influenced by a number of soil physical and chemical variables.
The amount of adsorbing substrate (iron and aluminum hydrous oxides, clay), the soil organic content,
and pH are probably the most important soil attributes. Hydrous oxide coatings of iron and aluminum,
38
-------
precipitated as amorphous or poorly crystalline material on particle surfaces, provide the principal
adsorbing substrate in most soils, and sulfate adsorption has been correlated to the abundance of
hydrous oxides (e.g., Chao et al., 1964b; Johnson and Todd, 1983; Fuller et al., 1985). Sulfate and
other anions can also be adsorbed at positive charge sites on edges of phyllosilicates. Adsorbed
sulfate has been correlated to abundance of clay in soils (e.g., Neller, 1959; Chao et al., 1962b;
Johnson et al., 1980), although clays are usually regarded as a secondary substrate for sulfate
adsorption (Johnson and Todd, 1983).
In recent years considerable attention has been given to the effects of soil organic matter on
anion adsorption. Several studies have recently demonstrated that organic matter can reduce, or
block adsorption of sulfate or other anions by soils (e.g., Barrow, 1967; Haque and Walmsley, 1973;
Gobran and Nilsson, 1988; Shaffer, 1989). Fitzgerald and Johnson (1982) hypothesized that blocking
results from competition for anion adsorption sites, with preferential sorption of fulvic acids, a
hypothesis supported by experimental results of Shaffer (1989) for a Virginia forest soil.
Along with the amounts of adsorbing substrates and of competing anions, pH provides a strong,
but indirect, control on sulfate adsorption by soils. Chao et al. (1964b) initially demonstrated that
sulfate adsorption increased as pH was reduced, and several subsequent studies (e.g., Hingston et
al., 1967; Couto et al., 1979; Nodvin et al., 1986) have shown similar pH effects. The effects of pH on
adsorption occur principally as a result of pH-dependent changes in the surface charge on inorganic
oxides; as pH is reduced, the ratio of OH2 to OH' ligands is reduced, resulting in increased positive
surface charge and anion adsorption capacity. Reductions in pH also decrease the dissociation of
organic acids (Stevenson, 1982), reducing the blocking effects of organic matter on sulfate adsorption.
The kinetics of sorption have generally been reported to be very fast; under experimental
conditions, soil solution sulfate concentrations have approached dynamic steady state within 5 to 15
minutes after addition of sulfate. The rapid approach to steady state is not unexpected, given that
sorption is a surface reaction; chemistry models are usually formulated to treat sorption as an
equilibrium process, rather than a kinetically constrained one. A second issue is reversibility of
sorption, especially as it affects soil response for scenarios of reduced sulfur deposition. Reported
reversibility of sorption ranges from less than 10 percent in some tropical soils (even in the presence
of phosphate, which forms strong ligands and under most situations results in virtually quantitative
displacement of sulfate from soils) (Bornemisza and Llanos, 1967) to complete, rapid desorption
(Weaver et al., 1985; Sanders and Tinker, 1975). The reversibility of sorption appears to decrease with
aging of sulfate on the soil (decreased desorption with time), temperature (less desorption as
temperature increases), and characteristics of the adsorbing substrate. An integrated interpretation of
the fragments of evidence seems to be that substrate factors that are positively correlated with high
sorption tend to be associated with reduced desorbablilty, perhaps as a result of differences in the
nature of sorption and the number of ligands exchanged.
Several processes besides adsorption can retain sulfate within watersheds. Formation and
dissolution of aluminum hydroxy-sulfate minerals, sulfate reduction, and immobilization/mineralization
of sulfur in soil organic matter have all been cited as probable sulfate controls or sinks on a site-
specific basis and suggested as potentially important regional controls. All three processes have been
39
-------
discussed at length by Church et al. (1989; Section 3) and by Turner et al. (1990) and data suggest
that they are, at most, minor retention mechanisms in most watersheds in the DDRP regions. Under
conditions of low soil pH and high dissolved sulfate, concentrations of sulfate and aluminum can be
controlled by precipitation of an aluminum sulfate mineral phase such as jurbanite (AIOHSO4),
basaluminite (AI4(OH)10SO4), or alunite ((K,Na)AI3(OH)6(SO4)2) (Adams and Rawajfih, 1977;
Nordstrom, 1982; Khanna etal., 1987). Except in areas of very .high sulfur fluxes [e.g., deposition at
the Soiling Forest, FRG (Khanna et al., 1987) or internal sources at Goettingen, FRG (Weaver et al.,
1985)], soil solutions are generally reported to be undersaturated for mineral phases such as jurbanite
or alunite (e.g., Cozzarelli, 1986; Dahlgren et al., 1990); interpretations are based on indirect data
(saturation indices), however, and are not unequivocal.
In intermittently or permanently anaerobic soils, in wetlands, and in lake sediments, substantial
retention of sulfate can occur by sulfate reduction (Weider and Lang, 1988; Bayley et al., 1986;
Schindler et al., 1986b; Baker et al., 1986b). Retention occurs principally by dissimilatory reduction
and principal reduction products are hydrogen sulfide, organic sulfur, or metal sulfides (evident in the
soil by gleying or mottling). Because sulfur is quickly reoxidized to sulfate upon drying and reaeration
of soil (Bayley et al., 1986; Nyborg, 1978), reduction provides a long-term sink only in environments
that are permanently anaerobic. The potential for sulfate reduction in soils in DDRP M-APP
watersheds is discussed in Section 7.
The majority of total sulfur in most soils occurs in organic form, principally as ester sulfate
(R-O-SO3) or as reduced, carbon-bonded sulfur (Bettany et al., 1973; David et al., 1982; Schindler et
al., 1986a; Johnson et al., 1982). Several studies have shown extensive cycling between organic and
inorganic forms of sulfur and among organic sulfur pools, and have suggested that organic pools can
be a large net sink for sulfate (e.g., Fitzgerald et al., 1982; Swank et al., 1984; David et al, 1984;
Schindler et al., 1986a). Soil analyses at Coweeta, North Carolina, suggest that both sorption and
organic accumulation are important sulfur sinks, at least in upper soil horizons (Strickland and
Fitzgerald, 1984; Strickland et al., 1987; Fitzgerald and Watwood, 1987). In contrast, in the
northeastern United States, analyses of sulfur transformation kinetics and sulfur isotopic ratios (Fuller
et al., 1986a,b) and sulfur input-output budgets (Rochelle and Church, 1987) suggest that systems are
near steady state and that net retention of organic sulfur, if any, is small.
3.3.2 Other Anlons
Along with sulfate, other acid anions contribute to mobile anion leaching in proportion to their
flux from watersheds. Leaching of nitrate has been noted as a potential source of acid transport to
surface waters, especially during snowmelt episodes (Galloway et al., 1987; Driscoll et al., 1987a,b;
Wigington et al., 1990). Inasmuch as the DDRP focuses on chronic rather than episodic acidification,
snowmelt effects are not considered here. In most forest systems in the eastern United States,
nitrogen is a limiting nutrient and is almost quantitatively retained by biological uptake and accretion in
biomass (Likens et al., 1977; Swank and Crossley, 1988), but recent studies suggest that nitrogen
throughput can be a significant contributor to chronic acidification at selected sites with mature forests
and/or very high rates of nitrogen deposition (C. Driscoll, pers. comm.). Evidence for such chronic
effects was not available when the DDRP was initiated and models and data for a regional assessment
40
-------
of nitrogen dynamics are not available, so changes in nitrogen dynamics have not been considered in
this Project. The role of nitrogen in episodic and chronic acidification is the topic of other ongoing
and proposed studies by the EPA (Wigington et al., 1990).
Organic acids also contribute significantly to the acid-base status of some surface waters. Krug
and Frink (1983) discussed the importance of natural soil acidification processes, and hypothesized
that acidic deposition could lower the pH of soils proximate to surface waters, decreasing the
dissociation of humic acids and decreasing the mobility of organic acids as mobile anions. They
suggested that the effect of acidic deposition has not been to acidify surface waters, but rather to
convert natural, organically acidic surface waters to sulfate-dominated waters through the decreased
mobility of organic acids. LaZerte and Dillon (1984) have presented evidence that the Krug and Frink
hypothesis is not supported for acidic lakes in Ontario, Canada, but inorganic-organic acid interactions
clearly can affect the pH and buffer capacity of surface waters. The potential dynamic effects of
organic acids on soil and surface water chemistry (as presented by Krug and Frink) have been
discussed by Turner et al. (1990) and considered in interpretation of NSWS surface water chemistry
(Baker et al., 1990). The potential for such effects was incorporated into the ILWAS model used for
DDRP projections in the NE and SBRP, but was not otherwise explicitly considered in DDRP analyses
for the NE and SBRP, or for the M-APP region. The probable importance of organic acids to lake or
stream acid/base status in most surface waters is low in all three regions considered by the DDRP,
particularly in the M-APP and the SBRP (Baker et al., 1990). The areal extent of organic soils is much
lower in the M-APP and the SBRP than in the NE (Section 5; Church et al., 1989, Section 5), and
dissolved organic carbon (DOC) concentrations in M-APP and SBRP stream waters are much lower
than in NE lakes. For streams in the M-APP target population, median and maximum DOC are only
1.1 and 2.0 mg/L, respectively (Kaufmann et al., 1988). Organic acids are a minor fraction of total
anion equivalents in stream water, and apparently play a very minor role in solution chemistry in
streams of the target population.
3.4 CATION SUPPLY PROCESSES
Mineral weathering and cation exchange are generally recognized as the dominant sources of
base cations to surface waters; both processes have been discussed at length by Church and Turner
(1986), Church et al. (1989), and Turner et al. (1990) and are only briefly summarized here. Along
with deposition inputs, mineral weathering represents the ultimate source of base cations in
watersheds. Base cations supplied from weathering are derived from a large pool, but are supplied at
a slow rate controlled by the kinetics of weathering reactions. Conversely, exchange is a surface
reaction involving a relatively small pool of cations held on organic material and on grain surfaces;
exchange reactions reach steady state on very short time frames and can be modelled as equilibrium
reactions.
Major rock-forming minerals are thermodynamically unstable in soil environments, and
weathering can be regarded as a unidirectional process. Weathering rates vary widely and depend on
the composition and structure of the component minerals [e.g., degree of polymerization of silica
tetrahedra, extent of Ca and Al substitution in feldspars (Huang, 1977)] as well as on the weathering
environment (Kittrick, 1977; Huang, 1977). Weathering rates under field conditions are poorly known,
41
-------
and have for the most part been estimated from cation budget data; the effects on in situ weathering
rates of environmental factors such as temperature and pH are virtually unknown. Laboratory data
indicate that weathering rates of major rock-forming silicate minerals increase (1) as a function of
hydrogen ion activity (raised to a fractional order typically between 0.2 and 0.5) (e.g., Wollast, 1967;
Helgeson et al., 1984; Chou and Wollast, 1985, Holdren and Speyer, 1985) and (2) with temperature
(Lasaga, 1981), although almost no data exist to characterize rate dependency at environmentally
relevant temperatures. Effects of organic ligands on weathering rates are uncertain (e.g., Huang,
1977; Mast and Drever, 1987).
Most watersheds in the DDRP M-APP sample are underlain by, and are mantled with soils
formed in situ from, clastic sedimentary rocks that are the product of one or more previous weathering
episodes. As the remnants of previous weathering episodes, these parent materials in most cases
consist predominantly of minerals that weather only slowly, with the result that weathering rates on
many watersheds in the M-APP region might be slow in relation to rates of sulfur deposition.
Cation exchange processes stabilize soil solution chemistry in two important respects; they
provide a buffer to maintain relative concentrations of cations in the soil solution, and they can also be
a net source or sink for basic or acidic cations. Cation exchange reactions are usually described in
terms of equilibria between pairs of species for aqueous and solid (i.e., exchangeable) phases (e.g.,
for calcium and aluminum):
= 3Ca
2+
2 AIX
where CaX2 and AIX3 represent cations held on the solid phase and Ca2+ and AI3+ are aqueous
species. For a given pair of ions, equilibrium between species can be described in terms of a mass
action equation by a selectivity coefficient, Kex :
3+
2+ 3
Kex = {AI [XCa]3 /{Ca2+ }
where the species in braces {} are expressed as the activity of aqueous species, and brackets [X] are
mole fractions of the solid phase species.
Under steady-state conditions, the cation exchange pool represents a dynamic steady state
between rates of cation input to the soil (deposition, decomposition of litter, mineral weathering) and
output from the soil (biological uptake and solution leaching). Under the perturbed conditions of
increased H+ and mobile anion input due to acidic deposition, cation leaching rates increase, and the
exchange pool can become an important net source of base cations to solution while acting as a sink
for hydrogen and/or aluminum. Because the exchangeable base pool is not only finite but also
relatively small in many forest soils, increased base cation leaching can result in significant depletion
of the exchange pool, and eventually will lead to a new steady state, albeit with reduced soil base
saturation and less efficient replacement of acid cations in solution (e.g., Bloom and Grigal, 1985;
Cosby et al., 1985b; Reuss and Johnson, 1986).
42
-------
3.5 OTHER PROCESSES
The routing of water through a watershed plays a very important role in the determination of
surface water chemistry and in the response of soils and surface waters to acidic deposition (e.g.,
Chen et a!., 1984; Peters and Driscoll, 1987). The effects of hydrologic routing were reviewed by
Church and Turner (1986) at the outset of the DDRP and are not discussed in detail here. The in-
depth determination of hydrologic flowpaths and residence times of waters in individual DDRP
watersheds was not within the time, budgetary, or logistic scope of the Project. Apparent associations
between watershed hydrologic parameters (or their surrogates) and surface water chemistry were
evaluated for DDRP watersheds in the NE and SBRP (Church et al., 1989; Rochelle et al., 1989b), but
equivalent analyses were not performed for the M-APP watersheds. Assumptions concerning
flowpaths and their effects on analyses are discussed in Sections 9 and 10, and a description of the
hydrologic module of the MAGIC model is presented in Section 10.
During the last decade, several studies have demonstrated that processes occurring within lakes
(and lake sediments) can significantly affect chemical budgets of lake systems (e.g., Cook and
Schindler, 1983; Schindler, 1986; Kelly et al., 1987; Baker and Brezonik, 1988; Baker et al., 1986a;
Brezonik et al., 1987). Several processes, especially dissimilatory reduction of nitrate and sulfate in
anaerobic lake sediments, can generate significant amounts of alkalinity and concurrently sequester
sulfate and nitrate (Rudd et al., 1986a,b; Kelly et al., 1987). The importance of these processes to
alkalinity, sulfate, and nitrate budgets increases in proportion to the hydrologic residence time of lakes
(Baker et al., 1986b; Kelly et al., 1987). Due in large part to the short residence time of water in most
lakes characterized by the NSWS (Linthurst et al., 1986; Kanciruk et al., 1986), the effects of in-lake
processes on alkalinity and sulfate budgets for those systems, including NE lakes in the DDRP target
population, are estimated to be small (Shaffer and Church, 1989). DDRP analyses in the SBRP and
M-APP regions are limited to stream systems; in-lake processes are restricted to a few small
impoundments on a few of these systems, and their effects on ANC, sulfur, and nitrate budgets are
assumed to be negligible.
Land use and changes in land use and vegetation can affect both the chemistry of surface
waters and their response to acidic deposition. Wetlands can function as a sink for nitrate and sulfate
and as a source for alkalinity and organic acids (Bayley et al., 1986; Weider and Lang, 1988).
Biomass accretion represents a substantial sink for base cations and nutrients from the soil (Krug and
Frink, 1983; Johnson and Todd, 1987). Anthropogenic use and disturbance can affect soils in a
variety of ways (e.g., acid mine drainage, soil amendments, depletion of soil nutrients) that can have
either an acidifying or an alkalinizing effect on soils and runoff. Relationships between current land
use and surface water chemistry were assessed for NE and SBRP watersheds by Church et al. (1989,
Section 8) and have been summarized by Liegel et al. (1991). In assessing the probable effects of
land use on surface water chemistry in the M-APP region, it is important to recognize that wetlands
occur on only a small proportion of watersheds in the target population, and their extent on most of
those watersheds is small (Section 5.5). Watersheds with major, land disturbances, such as
urbanization or oil/gas wells, were not included in the original NSS sample for the region (Kaufmann et
al., 1988), and sites with evidence of acid mine drainage were dropped from the DDRP sample and
target populations (Section 5.2). Projections of the potential effects of acidic deposition on long-term
43
-------
surface water chemistry are performed principally as comparisons among deposition scenarios, rather
than as forecasts of the actual probability that a particular condition will occur (see discussion in
Section 10). The latter would require reliable knowledge of future changes in land use in the region of
concern, knowledge that is not available.
The regional assessments conducted in the DDRP analyses have focused on a limited suite of
processes known to play a major role in controlling surface water chemistry and to affect the rate and
extent of surface water chemical response to acidic deposition. Project analyses have been centered
on sulfate adsorption and base cation resupply, the two sets of processes initially identified by
Galloway et al. (1983) and the NAS Panel on Processes of Lake Acidification (NAS, 1984). Research
conducted since the onset of the DDRP has provided continued evidence of the primary role played
by these processes, particularly as processes affecting the rate of chronic acidification. The DDRP
analyses have additionally considered a wide range of other processes, through their inclusion in Level
II and III models, through specific analyses to address individual processes such as hydrologic contact
and in-lake alkalinity generation, and through a diverse set of Level I regression analyses.
44
-------
SECTION 4
PROJECT APPROACH
4.1 INTRODUCTION
As H.B.N. Hynes (1975) once noted, "We must not divorce the stream from its valley in our
thoughts at any time. If we do we lose touch with reality." Although surface waters can be affected by
acidic deposition originating from emissions many miles distant, the concept of the watershed as a
unit is critical in understanding current and future aquatic effects. Indeed, for drainage lake and
reservoir systems in the Northeast, Upper Midwest, and Southern Blue Ridge Province, most ANC
production occurs as a result of biogeochemical processes within the surrounding watershed (Section
7.2; Shaffer et al., 1988; Shaffer and Church, 1989).
Because of the importance of watershed processes (especially those occurring in soils) in
determining future aquatic effects, new data on these processes and on related soil pools and
capacities were required. Initially, we considered using existing regional soils data in the DDRP
analyses. Existing soils databases, however, have serious deficiencies with respect to the needs of
the Project. First, because of the economic importance of croplands, such data are available primarily
for lowland agricultural regions; surface water acidification, however, occurs principally in relatively
undisturbed upland systems. Second, such databases generally do not include chemical
characterizations of a number of key variables relevant to soil chemical interactions with acidic
deposition (e.g., sulfate adsorption capacity and unbuffered cation exchange capacity).
After considering these factors, we decided that the Project required a new regional soils
database, necessitating a major soil survey effort (Sections 5.1 - 5.5; also see Lee et al., 1989a;
Church, 1989). We further concluded that this survey should enable the specific soils (and specific
soil types) to be linked with the NSWS databases that describe the chemistry of low ANCJakes and
streams. Accordingly, we adopted the approach outlined in this section and illustrated in Figure 4-1.
4.2 SOIL SURVEY
4.2.1 Watershed Selection
The DDRP watersheds selected comprise a high-interest subset of lake and stream systems
surveyed in the NSWS. We selected an adequate number of watersheds to allow for (1) reasonably
broad regional coverage and (2) statistical examination of interrelationships (deposition watershed
characteristics:surface water chemistry) and model projections of response. Because these
watersheds are a subset of a probability sample (i.e., the National Stream Survey - see Kaufmann et
al., 1988), results can be extrapolated to a specified population of interest. Further details on
watershed selection are provided in Section 5.2 (also see Lee et al., 1989a). Regional population
estimation is discussed in Section 6.
45
-------
Project Design^'
1 '
*
Watershed Selection
i
*
Watershed Mapping
*
Development of Soil
Sampling Classes
i
Soil Sampling and
Field Measurements
i
Soil Prepraration
Chemical/Physical
Laboratory Analysis
I
*
Data Analysis
i
Supporting Regional
Datasets
I •
•
Database Management
]f Reporting?" /
Figure 4-1. Steps in the Direct/Delayed Response Project (DDRP) approach. Asterisks denote
steps that received significant support from Geographic Information Systems
(GlS)-based activities (Campbell and Church, 1989; Campbell et al., 1989).
46
-------
4.2.2 Watershed Mapping
The USDA Soil Conservation Service (SCS) prepared maps of soils, vegetation, land use, and
depth to bedrock for each DDRP watershed. Bedrock geology was obtained from existing state
geology maps. SCS mapping was at a scale of 1:24,000 and was at a "second order" intensity
(comparable to most county soil surveys). An important part of this mapping was the regional
correlation of map unit names and definitions, a common procedure at the county or state level, but a
much greater challenge at the regional scale of this Project. Additional maps of land use and
wetlands were developed by interpreting infrared stereo aerial photographs at a scale of 1:12,000, with
land use delineated to 2.5 ha and wetlands to 0.4 ha. Watershed mapping is discussed in detail in
Section 5.4.
4.2.3 Sample Class Definition
Because many soil components were mapped in the study regions (e.g., 280 in the M-APP),
characterizing each one physically and chemically was not feasible. Instead, sample classes were
defined for each region, and individual soils were assigned to those classes based on (1) expert
knowledge of the soils mapped and (2) expectations of the potential responses of those soils to acidic
deposition. Soils selected from these classes were sampled across the study regions. Soils were
aggregated within sampling classes to develop characterizations (e.g., class means and variances)
that were used to "rebuild" or represent (e.g., by mass or area weighting) the characteristics of study
watersheds. Details of the sample class selection are provided in Section 5.5.1 and by Lee et al.
(1989b).
4.2.4 Soil Sampling
We developed a procedure that allowed random selection of soil sampling sites within the
context of expert classification. This procedure was designed to ensure that adequate and complete
coverage was obtained of both the sampling classes and the watersheds across the regions. Details
are given in Sections 5.5.2 and 5.5.3.
4.2.5 Sample Analysis
Samples were analyzed by independent soil laboratories under contract through EPA's
Environmental Monitoring and Systems Laboratory - Las Vegas (EMSL-LV). A rigorous quality
assurance (QA) program was implemented to ensure the quality of these analyses. Sample analyses
are discussed in Section 5.5.4.
4.2.6 Database Management
Management of the soil survey databases involved operations at the Environmental Research
Laboratory - Corvallis (ERL-C), EMSL-LV, and Oak Ridge National Laboratory (ORNL). Centralized
database management was maintained at ORNL, with backup at ERL-C. Database management
activities in the DDRP are further discussed in Sections 5.4 and 5.5.
47
-------
4.3 OTHER REGIONAL DATASETS
Because of the regional nature of the Project, we required estimates of precipitation,
atmospheric deposition (wet and dry), and surface water runoff (as runoff depth), all of which were
generated in a standardized manner across the eastern United States. Study sites for the DDRP were
selected statistically, and most sites had no direct information for the above variables. Furthermore,
time and budgetary constraints precluded the instrumentation of sites and, thus, the direct acquisition
of such data. These estimates, therefore, had to be developed within the Project.
4.3.1 Atmospheric Deposition
Throughout the DDRP, the acquisition and development of internally consistent regional datasets
on atmospheric inputs was a challenging task. For the previous DDRP studies in the NE and SBRP,
two types of datasets were developed (Church et al., 1989). One dataset representing long-term
annual average (LTA) conditions was constructed for use in the correlative analyses in the Project
(Section 4.4.1). The temporal resolution of this dataset is annual. A second atmospheric deposition
dataset was constructed for use in the watershed modelling analyses of the Project (Section 4.4.3).
This dataset provides daily estimates of precipitation (to "drive" the hydrologic subroutines of the
watershed models) and monthly inputs of atmospheric deposition corresponding to a "typical year"
(TY).
For both datasets, precipitation amount and chemical concentrations were estimated from the
Acid Deposition System (ADS) network (Wampler and Olsen, 1987). Wet deposition was determined
as the product of these measures. Dry sulfur deposition was estimated from simulations using the
Regional Acid Deposition Model (RADM) (R. Dennis and S. Seilkop, pers. comm.; Clark et al., 1989).
Full estimates of dry deposition (including both fine and coarse particles) for other ions were not
directly available from any source at the time the Project was implemented, and had to be developed
within, or in association with, the Project (see Eder and Dennis, 1990). Church et al. (1989) described
the development of these two datasets.
In our Level I and II analyses (see Section 4.4) for the NE and SBRP, we found that application
of the TY and LTA deposition data yielded very similar results. Therefore, we did not repeat the
generation of these two types of deposition datasets for our analyses in the M-APP. Inasmuch as TY
deposition was needed for the Level III modelling analyses, we generated the TY dataset only (Section
5.6) and used it as the base case in Level II analyses.
4.3.2 Runoff Depth
Because direct runoff measurements were lacking for the selected watersheds, we relied upon
regional maps of annual runoff depth. Investigation of the maps available at the start of the Project
yielded no single map with a resolution finer than 5 inches of runoff depth. We therefore enlisted the
U.S. Geological Survey (Madison, Wisconsin) to produce an annual runoff map for the period 1951-80
(Krug et al., 1990), corresponding to long-term precipitation records used to estimate deposition. As
part of this work, we performed a quantitative uncertainty analysis of estimates of long-term runoff from
48
-------
the Krug et al. map (Rochelle et al., 1989b). More information on the development and application of
these runoff data within the Project are given in Section 5.7.
4.4 DATA ANALYSIS
A variety of analyses have been undertaken within the Project. Many analyses were performed
by the EPA and contractor staff at ERL-C. Others were performed by extramural cooperators in close
coordination with ERL-C staff. Data analyses within the Project are classified into three "levels",
according to the complexity of the analyses and the degree of reliance upon knowledge, or
hypotheses, of process interaction within watersheds. For example, Level I analyses presuppose the
least about our knowledge of the way watersheds "operate", whereas Level III analyses depend upon
more comprehensive knowledge of system behavior.
4.4.1 Level I Analyses
Level I analyses for the NE and SBRP included constituent input-output budget estimates and
statistical analyses (Church et al., 1989). The leaching of the mobile anion sulfate is considered to be
a key process in long-term acidification. Accordingly, one part of the Level I analyses was to
determine the retention of atmospherically deposited sulfur within watersheds. We examined annual
watershed input-output budgets for sulfur, based on detailed studies at a few sites and relatively
sparse data from many sites (Church et al., 1989). For the sake of completeness in this report, we
repeat parts of those analyses (Section 7) with special emphasis on the M-APP Region.
For the NE and SBRP regions, the other part of Level I analyses was the statistical evaluation of
interrelationships among atmospheric deposition, mapped watershed characteristics, soil chemistry,
and current surface water chemistry (e.g., see Rochelle et al., 1989a). One goal of this evaluation was
to verify that the processes and relationships incorporated in the Level II and III analyses reasonably
represent the systems under study. To a reasonable degree, that goal was accomplished (see
Section 8 of this report), and for reasons of economy of time and of effort, we chose not to repeat
those kinds of analyses for the M-APP.
4.4.2 Level II Analyses
The Level II analyses use relatively restricted models of key processes that regulate the
dynamics of (1) base cation supply and (2) watershed retention of atmospherically deposited sulfur.
The models are used to project how these processes might affect conditions in the DDRP watersheds,
and in the surface waters that drain them, under continuing or altered future levels of atmospheric
sulfur deposition. The models used to investigate and project base cation supply are the Bloom-Grigal
model (Bloom and Grigal, 1985) and the Reuss-Johnson model (Reuss and Johnson, 1985, 1986).
We modelled watershed retention of sulfur as sulfate adsorption, according to the approach presented
by Cosby et al. (1986b). The models are run independently of one another and of other watershed
factors, such as forest accretion, that might affect watershed response. The analyses and results are
given in Section 9.
49
-------
4.4.3 Level III Analyses
In the DDRP Level III analyses, integrated watershed models are used to project future effects of
atmospheric sulfur deposition on surface water chemistry. Three models specifically developed to
investigate the effects of acidic deposition on watersheds and surface waters were applied in our work
for the NE and SBRP (Church et al., 1989): (1) the Model for Acidification of Groundwater in
Catchments (MAGIC) (Cosby et al., 1985a,b,c; Cosby et al., 1986a,b), (2) the Enhanced Trickle Down
(ETD) Model (Lee, 1987; Nikolaidis et al., 1988; Schnoor et al., 1986), and (3) the Integrated
Lake-Watershed Acidification Study (ILWAS) Model (Chen et al., 1983; Gherini et al., 1985). These
three models were selected on a competitive, externally peer-reviewed basis via EPA's standard
Cooperative Agreement funding mechanism. The sequence followed included a public announcement
of the Request for Proposals, committee review of pre-proposals, and external peer review of full
proposals.
The three models were run by each of the respective groups that developed them using
common DDRP datasets for forcing functions (e.g., rainfall, runoff, atmospheric deposition) and state
variables (e.g., soil physical and chemical variables). The modellers made projections of changes in
annual average surface water chemistry for DDRP watersheds in the Northeast (NE) and the Southern
Blue Ridge Province (SBRP) for at least 50 years for two scenarios of atmospheric sulfur deposition:
(1) continued deposition at current levels (for both regions) and (2) altered deposition over the next 50
years (i.e., a decrease in the NE and an increase in the SBRP).
Although projections of the potential effects of continued sulfur deposition were variable
(occasionally highly variable) among the three models for any individual watershed (Church et al.,
1989), projections among the models were extremely similar on a regional basis. Regional analyses
and projections are, of course, the principal focus of the DDRP. Rather than repeat very lengthy, time
consuming, and expensive analyses with three separate watershed models for the M-APP Region, we
applied the singfe model that, among the three models as exercised by their developers in the NE and
SBRP analyses, provided the most complete, timely, and economical calibrations and simulations of
effects on watersheds and receiving waters. This was the MAGIC model. The model was applied by
contractor staff on site at ERL-C using the same procedures that were used for modelling analyses in
the NE and SBRP. Close cooperation with Dr. J. Cosby (Duke University), the developer of MAGIC,
Dr. P. Ryan (Oak Ridge National Laboratory), who participated in the NE and SBRP analyses and who
has a very high degree of familiarity with the model, and Dr. K. Thornton (FTN & Associates), who
coordinated the Level III analyses for the NE and SBRP, insured that application of MAGIC for M-APP
watersheds paralleled the application for the other regions. We present results of the application of
MAGIC to the DDRP watersheds in the M-APP Region in Section 10 of this report.
4.4.4 Integration of Results
As for the DDRP analyses in the NE and SBRP, de facto integration of interim results for the
M-APP has taken place during the course of the analyses for this region. As noted in Section 2, the
principal bottom line of the DDRP (i.e., time dynamic projections of the long-term effects of sulfur
deposition on regional surface water chemistry) comes from the dynamic watershed simulations
50
-------
performed in the Level III analyses. In Section 11 of this report, we discuss the manner in which the
results from the previous Level I and current Level II analyses support and expand upon the Level III
findings.
51
-------
-------
SECTION 5
DATA SOURCES AND DESCRIPTIONS
5.1 INTRODUCTION
Section 5 presents enough information concerning the design of the Project and data acquisition
within the DDRP to familiarize the reader with the characteristics of the region studied and to allow the
reader to evaluate the analyses performed in Sections 7 through 10. Many data have been generated
and used by the Project during its course. Although a complete listing of the data is not presented
here, descriptions of the way the data were gathered within the Project or obtained from-other sources
are presented, along with pertinent examples or summaries of the data.
5.2 STUDY SITE SELECTION
In selecting study sites, the intent was to focus on regions with watersheds potentially sensitive
to acidic deposition (Section 2.3), but exhibiting a wide contrast in both soil and watershed
characteristics and levels of deposition.
5.2.1 Site Selection Procedures
The procedures for selecting the DDRP sample watersheds for the Mid-Appalachian Region
(M-APP) differed somewhat from those used to select DDRP sites in the NE and SBRP, based on
experience gained during the course of the project. The DDRP site selection builds on the design of
the National Stream Survey (NSS), which is discussed below. Kaufmann et al. (1988) have given a
complete description of the NSS design.
5.2.2 National Stream Survey Design
The NSS framed the target population by defining a stream reach as the length of a "blue-line"
stream, on USGS 1:250,000-scale maps, that lies between the downstream and upstream confluences
with other blue-line streams, or the upper stream boundary if no upper confluence is present (Figure
5-1).
A two-stage sampling scheme was used for stream selection. In the first stage, an area/point
frame was used to select the sample of stream reaches. A rectangular grid of points, separated by a
scaled distance of approximately 13 km (8 mi), was positioned at random over a 1:250,000
topographic map. The first stream reach intersected by a line from each point drawn downslope
perpendicular to the contour lines was included in the first-stage sample (Figure 5-1). If the reach
extended beyond the study region, if the watershed was larger than 155 km^, if the reach had a
Strahler order over 5, if over 20 percent of the watershed was in an urban area, or if the reach drained
into a reservoir, the reach was designated as noninterest and dropped from the sample. The inclusion
probability for a reach in the target population was proportional to the watershed area draining directly
into the reach. A separate stratum of low ANC sites was identified (areas where ANC was expected to
52
-------
Non-Headwater Reach
Headwater Reach
Figure 5-1. Representation of the point frame sampling procedure for selecting NSS Stage I
reaches. Area a1 represents the direct drainage area to the lower node of non-
headwater reaches, or the total drainage area to the lower node of headwater
reaches. Area a2 is the total drainage area to the upper node of non-headwater
reaches.
53
-------
be below 50 ^eq/L), in order to increase resolution in the high-interest part of the target population
(Kaufmann et al., 1988).
In the second stage, variable probability sampling was used to select a subsample of these
reaches. All sites in the low ANC stratum were automatically included. Also, some sites had such
small drainage areas that they entered the Stage II sample with inclusion probabilities of one. Except
for these cases, reaches were selected with inclusion probabilities inversely proportional to the direct
drainage areas of the stream reaches.
Some of the selected reaches were considered to be noninterest sites on the basis of chemistry,
field observations, and watershed characteristics. The exclusion criteria include a lack of stream
channel, a lack of water in the stream, existence of acid mine drainage and industrial pollution,
existence of tidal influence, and effects due to precipitation events at or near the time of sampling.
5.2.3 DDRP Site Selection
The M-APP DDRP data were obtained from 36 NSS stream watersheds. These watersheds were
selected from the NSS dataset in an iterative procedure described below.
In the first step, watersheds in only subregions 1D, 2B, and 2C were screened. Sites in these
subregions were eliminated if they had (1) ANC > 400 ^aeq/L, (2) watershed area of more than 3,000
ha, as defined based on the downstream sampling node, (3) chemical data available for only one of
the two nodes of the stream reach, or (4) observed residual effects of acid mine drainage from old
spoil banks within the watershed. (We used a criterion of 3,000 ha, for watershed size to be
consistent with procedures for the NE and SBRP regions [Church et al., 1989]. Mapping of
watersheds above this size becomes prohibitively difficult and time consuming at the scales we used.)
Our screening produced an initial list of 79 possible watersheds. Aerial photography was obtained for
these sites in preparation for soil mapping activities.
In the next step, watersheds with ANC > 200 fieq/L were excluded. These watersheds were not
considered to be at as high a risk of future acidification as systems with lower ANC. This reduction
helped bring the total budget for the M-APP survey in line with available funds.
The topo sheets of all remaining watersheds were checked for mine spoils or other
disturbances, and watersheds with such symbols were excluded. Also, some watersheds in subregion
1D were excluded because they were in the region already mapped in the DDRP Northeast Soil
Survey (Church et al., 1989). These watersheds were close to the line separating Major Land
Resource Areas R140 (Glaciated Allegheny Plateau and Catskill Mountains) and S147 (Northern
Appalachian Ridge and Valley). One further watershed was excluded as disturbed during field
observation by the Soil Conservation Service (SCS), who were cooperators in the DDRP. This
reduced the sample to 36 watersheds. Table 5-1 provides basic physical and chemical information on
the stream reaches studied, including sampling weights and inclusion probabilities. Tables 5-2 and
5-3 and Plate 5-1 further identify and locate the stream watersheds.
54
-------
Table 5-1. S'tream Identification (ID), Weight, and Inclusion Probabilities for the M-APP
DDRP Sample Watersheds
Stream ID
1D029023
1D029031
1D029042
1D029043
1D036011
1D036017
1D037005
2B036028
2B036046
2B036062
2B041008
2B047032
2B047036
2B047066
2B047076
2B047089
2C028069
2C028070
2C028075
2C029002
2C029016
2C029020
2C035027
2C041002
2C041039
2C041040
2C041045
2C041051
2C046005
2C046033
2C046034
2C046041
2C046050
2C047007
2C047010
2C057004
ANC
(weq/L)
32.3
61.3
13.2
1.5
13.3
102.0
4.1
135.3
62.7
199.6
139.4
93.5
75.5
168.5
5.0
66.8
-3.3
9.4
2.8
166.9
26.7
9.8
109.2
83.5
76.3
48.5
144.0
-9.6
53.8
6.2
4.6
155.7
86.8
91.8
27.6
16.0
Watershed
Area
(km2)
11.6
10.4
16.5
17.0
5.6
9.6
10.9
3.7
10.2
18.7
16.3
10.4
23.2
1.4
3.5
17.3
4.7
17.0
11.0
10.7
5.7
23.6
19.8
0.98
5.8
1.8
19.8
4.3
0.73
4.4
1.6
2.3
0.52
3.4
8.7
3.2
Elevation
(m)
439
333
363
547
281
333
183
122
192
271
282
823
462
165
354
427
518
414
428
497
509
344
564
500
576
658
485
558
314
704
879
677
603
607
920
721
Reach
Length
(km)
6.6
7.0
8.0
6.5
3.5
5.8
5.3
3.2
6.0
1.9
0.8
2.6
2.7
2.2
4.1
1.6
3.3
6.6
6.1
4.6
5.8
9.3
8.6
1.4
1.1
2.0
4.8
3.3
1.2
0.74
1.7
3.1
1.1
4.6
1.5
1.5
Sampling
Weight
59.74
59.74
10.03
9.76
59.74
59.74
15.20
317.80
317.80
317.80
317.80
317.80
317.80
317.80
47.41
58.72
35.56
9.77
15.13
145.26
29.09
7.03
145.26
168.42
145.26
145.26
145.26
145.26
228.57
320.00
145.26
145.26
320.00
145.26
304.76
145.26
Inclusion
Probability
0.01674
0.01674
0.09970
0.10246
0.01674
0.01674
0.06579
0.00315
0.00315
0.00315
0.00315
0.00315
0.00315
0.00315
0.02109
0.01703
0.02812
0.10235
0.06609
0.00688
0.03438
0.14225
0.00688
0.00594
0.00688
0.00688
0.00688
0.00688
0.00438
0.00312
0.00688
0.00688
0.00312
0.00688
0.00328
0.00688
55
-------
Table 5-2. Stream Identification (ID) and Name, and State and Latitudinal/Longitudinal
Location of the M-APP Sample Watersheds, Sorted by Stream ID
Stream ID
1D029023
1 D029031
1 D029042
1D029043
1D036011
1D036017
1D037005
2B036028
2B036046
2B036062
2B041008
2B047032
2B047036
2B047066
2B047076
2B047089
2C028069
2C028070
2C028075
2C029002
2C029016
2C029020
2C035027
2C041002
2C041039
2C041040
2C041045
2C041051
2C046005
2C046033
2C046034
2C046041
2C046050
2C047007
2C047010
2C057004
Stream Name
North Branch Rock Run
East Branch Wallis Run
Heberly Run
South Branch Bowman Creek
Stony Run
Locust Creek
Jeans Run
No Name
Lower Little Swatara Creek
Burns Creek
Piney Creek
Elk Run
Bible Run
No Name
Lower Lewis Run
North Fork Moormans River
Whitney Run
Coldstream Run
Bear Run
Upper Dry Hollow
East Branch Big Run
Wolf Run
Williams Run
Fulton Run
- Buffalo Creek
Thunderstruck Creek
Right Fork Clover Run
Coal Run
No Name
Johnson Run
Hateful Run
No Name
Hedricks Creek
Crawford Run
Clubhouse Run
Butler Branch
State
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
WV
VA
VA
VA
VA
PA
PA
PA
PA
PA
PA
PA
PA
WV
WV
WV
WV
WV
WV
WV
WV
WV
WV
WV
WV
Latitude
0 »
41
41
41
41
40
40
40
40
40
40
39
38
38
38
38
38
41
41
41
41
41
41
40
39
39
39
39
39
38
38
38
38
38
38
38
37
32
26
19
20
52
47
52
37
33
13
47
37
38
24
18
9
7
6
0
40
11
5
28
56
15
14
8
2
41
20
21
13
7
45
37
57
II
48
17
35
55
55
20
22
26
8
40
37
55
40
19
18
55
42
46
54
49
46
23
51
34
39
55
51
23
8
50
4
16
31
32
56
22
Longitude
O » II
76
76
76
76
76
76
75
76
76
77
78
79
79
78
78
78
78
78
78
77
77
77
78
79
79
79
79
79
80
80
80
80
80
79
79
80
52
51
20
12
14
3
46
57
16
42
24
35
0
16
44
44
35
23
34
46
44
52
46
25
45
36
42
36
33
24
15
42
58
55
45
56
16
40
49
24
4
52
24
8
5
55
18
9
40
41
45
41
55
41
58
31
43
7
18
57
18
3
54
59
8
29
31
55
56
24
37
34
56
-------
Table 5-3. Stream Identification (ID) and Name, Sorted by State - M-APP
Sample Watersheds
State
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
PA
VA
VA
VA
VA
WV
WV
WV
WV
WV
WV
WV
WV
WV
WV
WV
WV
WV
Stream ID
1D029023
1D029031
1D029042
1D029043
1D036011
1D036017
1D037005
2B036028
2B036046
2B036062
2B041008
2C028069
2C028070
2C028075
2C029002
2C029016
2C029020
2C035027
2C041002
2B047036
2B047066
2B047076
2B047089
2B047032
2C041039
2C041040
2C041045
2C041051
2C046005
2C046033
2C046034
2C046041
2C046050
2C047007
2C047010
2C057004
Stream Name
North Branch Rock Run
East Branch Wallis Run
Heberiy Run
South Branch Bowman Creek
Stony Run
Locust Creek
Jeans Run
No Name
Lower Little Swatara Creek
Burns Creek
Piney Creek
Whitney Run
Coldstream Run
Bear Run
Upper Dry Hollow
East Branch Big Run
Wolf Run
Williams Run
Fulton Run
Bible Run
No Name
Lower Lewis Run
North Fork Moormans River
Elk Run
Buffalo Creek
Thunderstruck Creek
Right Fork Clover Run
Coal Run
No Name
Johnson Run
Hateful Run
No Name
Hedricks Creek
Crawford Run
Clubhouse Run
Butler Branch
57
-------
Plate 5-1. ANC at lower nodes of DDRP stream reaches in the M-APP Region.
58
-------
-------
MID-APPALACHIAN REGION
DDRP Study Site Locations
Stream ANC
ANC (ueq/L)
100
-APP Study Area
100 - 200
'„- -E r.r-- ll ,•>#*\ iJ*r,<^'M
\- l - '""J r"^"'^ A •" ->----"-
, ^ "„,, "-r,, *---"* - ',iC\ --, ,-,-' ,-
-------
-------
5.2.4 Final M-APP DDRP Target Population
The final DDRP target population for the M-APP represents 5,496 stream reaches (15,239 stream
km) based on a sample of 36 NSS watersheds that satisfied the DDRP selection criteria. This is a
subset of the NSS target population of stream reaches (25,715 stream reaches, 69,569 km) in NSS
subregions 1D, 2Cn, and 2Bn (Kaufmann et al., 1988). The M-APP target population represents
relatively undisturbed stream reaches with watershed areas less than 3,000 ha and ANC values less
than 200 ^ueq/L
5.3 DATA FOR NSS STREAM REACHES IN THE MID-APPALACHIAN REGION
5.3.1 Spring Baseflow Index Sampling
The index sampling concept for Phase I of the NSS is described in the following paragraphs.
"Like the ELS-l components of the NSWS, the NSS-I relied on samples taken during
an appropriate season from a representative sample of water bodies to provide an index
of the chemical characteristics of the target population (Messer et al., 1986). In the
Eastern and Western Lake Surveys (Linthurst et al., 1986; Landers et al., 1987), a single
mid-lake sample taken during well-mixed conditions at fall turnover provided a reasonably
good spatial representation of the nonlittoral lake water volume. Furthermore, this fall index
sample for lakes can be related to water quality during other seasons of the year when
chemical conditions may be more critical for biota (Driscoll and Newton, 1985; Newell,
1987). In lakes, relatively long hydraulic residence times (low flushing rates) tend to
integrate the inputs of water and dissolved materials from the lake watershed, which
reduces that portion of the chemical variability caused by changes in input rates. Streams
generally exhibit greater within- and among-season variability than do lakes. Since streams
have little temporal integrative capacity within their channels, it is necessary to draw an
index sample during a period of the year that is expected°to exhibit chemical
characteristics most closely linked to acidic deposition or to its most deleterious effects.
"Sampling the relatively stable chemistry of late summer baseflows dominated by
groundwater, for example, would provide a poor index of potentially limiting conditions
during winter and spring periods when the stream water is poorly buffered against pH
changes. The choice of the spring index sampling period for streams was based on a
literature search followed by a series of meetings with hydrologists, biochemists, and
fishery experts in Pennsylvania, Virginia, North Carolina, Florida, and Arkansas to discuss
ongoing projects involving stream chemistry and fisheries in the proposed NSS-I study
areas (U.S. EPA, 1984). The choice involved a trade-off between minimizing within-season
and episodic chemical variability and maximizing the probability of sampling during
chemical conditions potentially limiting for aquatic organisms.
"A number of sources of stream chemistry data from several geographic areas
support the choice of a spring index sampling period for observing prolonged periods of
low pH and ANC. Ford et al. (1986), for example, summarized the results of four recent
(1984-1985) studies of seasonal and short-term variability in six second- and third-order
streams in the Catskill Mountains of New York (Murdoch, 1986), the Laurel Hills of
Pennsylvania (Witt and Barker, 1986), the Southern Blue Ridge Province of North Carolina
and Tennessee (Olem, 1986), and the Ouachita Mountains of Arkansas (Nix et al., 1986).
Minimum flow-weighted pH values and concentrations of base cations and ANC occurred
during the spring at almost all sites. Those sites with minimum values during the winter
had spring values nearly as low.
59
-------
"For a spring index sampling period to be biologically relevant, however, sensitive
life-stages of aquatic biota must also be present during the sampling period. Studies have
indicated that all life stages of fish are not equally sensitive to acidity and chemical
constituents that accompany low pH conditions in surface waters. Some of these studies
involved observations of acidic lakes and streams in which viable eggs were found
together with older age classes of fish that appeared to be spawning successfully, but in
which young age classes were absent (e.g., Beamish et al., 1975; Muniz and Leivestadt,
1980; Kelso and Gunn, 1982; Gunn and Keller, 1984; Sharpe et al., 1984). Such a
population structure suggests more pronounced effects of acidity on larval fish than on
egg hatching or adult survival. These field observations are in agreement with laboratory
bioassays that also indicate greater sensitivity of fry to low pH conditions, relative to other
fish life stages (Schofield, 1976; Haines, 1981). Fry of the most important sport fish are
present in the NSS-I study area during the March 15 - May 15 period. Fry of some trout
(Salmo spp.) populations may also be present at other times of the year.
"In summary, spring appears to be the most appropriate index sampling period for
streams, because ANC is typically low, and life stages of aquatic biota that are sensitive to
low pH are likely to be present at this time. The low ANC during the season minimizes
buffering against episodic pH changes accompanying high runoff. Although pH and ANC
depressions can also occur during other seasons, they may be more pronounced during
the spring because short hydraulic residence times in the soil during the spring minimize
acid neutralization. Also, acid-sensitive, swim-up fry of key fish species are typically
present in streams during the spring in many parts of the United States. The index
sampling period for the NSS-I thus was chosen as the time period following snowmelt but
prior to leafout (mid-March to mid-May, depending on the subregion). Results of the NSS-I
Pilot Survey in the Southern Blue Ridge showed very little difference in separate
population distributions of pH, ANC, and major cations and anions based on three
successive spring baseflow samples during this sampling window (Messer et al., 1986,
1988). The occurrence of large episodic chemical changes over the course of hours or
days during storm runoff, however, makes the use of spring samples for indexing water
chemistry difficult, unless sampling during such events is avoided (Messer et al., 1986). To
avoid alterations in index chemistry caused by atypical stormflow samples, the NSS-I
avoided sampling within 24 hours following significant rain events (> 0.2 inches).
"Unlike lakes, for which a single mid-lake sample taken during well-mixed conditions
at fall turnover can provide a reasonably good spatial representation of the nonlittoral
lakewater volume, a sample taken at a single point on a stream reach would not
adequately describe chemistry for the whole length of the reach (Messer et al., 1986).
Streams were expected to exhibit substantial trends in chemistry over their length at any
given time during the spring index period. To incorporate this variability and to establish a
basis for quantifying relationships between upstream and downstream chemistry on
sample reaches, samples from both ends of the reaches were collected in the NSS-I."
(From Kaufmann et al., 1988; Section 2.5, "Index Sampling".)
As discussed in 5.2.3, DDRP study watersheds in the M-APP were defined based upon the
downstream nodes of the reaches sampled.
5.3.2 Chemistry of DDRP Stream Reaches
The complete chemistry at the downstream nodes of the stream reaches used to define the
DDRP study watersheds in the M-APP has been given by Messer et al. (1986) and Sale et al. (1988)
and will not be repeated here. Figure 5-2 gives the pH-ANC relationship for samples taken at the
downstream nodes of the NSS reaches in the relevant subregions for ANC < 200 ^eq/L The ANC
60
-------
8
7-
6-
5-
o
o
r°°°o
o o o o
o
oo o
•cP
o
•
o
4
-100
o 100
ANC (weq/L)
200
Figure 5-2. The pH-ANC relationship for samples with ANC < 200 fieq/L taken at the
downstream nodes of stream reaches in the NSS in subregions 1D, 2B, and 2C.
Shown are the relationships for all such samples from the NSS and samples for
the downstream nodes of DDRP study reaches in the M-APP. DDRP reaches are
denoted by filled circles, and non-DDRP reaches of the NSS are denoted by open
circles. Samples are the average of two samples, with samples taken during
precipitation events excluded.
61
-------
referenced for the stream reaches is the modified Gran ANC. Table 5-4 provides a comparison of
ANC reference values for the target populations of NSS Phase I (Subregions 1D, 2Bn, and 2Cn) and
M-APP DDRP, respectively.
5.4 MAPPING PROCEDURES AND DATABASES
The first step in gathering the terrestrial information required to characterize the study
watersheds was to map the watershed attributes. This mapping was designed to include all the major
characteristics thought to be important in determining the response of surface waters to acidic
deposition for watersheds selected to represent the study region. Existing terrestrial databases were
examined and found to be inadequate for the needs of this assessment (Lee et al., 1989a).
Specific resource inventories of soils, depth to bedrock, drainage, forest cover type, and land
use were designed within the Project and implemented through the assistance of the USDA Soil
Conservation Service (SCS) in the states of Pennsylvania and West Virginia. Mapping began in
October 1987 for the M-APP region and was completed in August 1988. Sixteen field soil scientists
conducted the mapping of 36 watersheds, an area of 34,513 ha (85,248 acres).
Both color-infrared and black-and-white prints of aerial photographs at a scale of 1:24,000 were
obtained for the watersheds in this project to be used for field mapping. Map unit delineations of the
watershed attributes were drafted on mylar overlays at a scale of 1:24,000, digitized in separate layers,
and entered into a Geographic Information System (GIS). Film positives of USGS 7.5-minute
topographic quadrangles were used as rectified base maps; photographically enlarged 15-minute
quadrangles were used where 7.5-minute maps were not available. The soil map legend, map unit
composition, characteristics of the map unit components, and soil transect data collected during the
field mapping were entered into an interactive microcomputer data management system.
Performance and direction of mapping activities were modelled after the organization of the
National Cooperative Soil Survey (Soil Survey Staff, 1983). The Mapping Task Leader for the DDRP,
located at the U.S. EPA Environmental Research Laboratory in Corvallis, Oregon (ERL-C), had overall
responsibility for mapping and coordinated all mapping activities. A Regional Coordinator/Correlator
(RCC), an independent contractor, provided quality assurance/quality control (QA/QC) for the field
mapping. The RCC provided training in and adherence to mapping protocols to ensure regional
consistency, maintained a uniform regional mapping legend, participated in at least one field review of
the mapping in each state, and evaluated mapping activities to assure quality. Mapping quality was
assessed by field checking the mapping on 20 percent of the watersheds and conducting transects on
selected map units.
The State Soil Scientist (USDA-SCS) in each state, with the support of soil scientists on the State
Soils Staff, was responsible for mapping activities in that state. This included supervising and coordin-
ating field crews, performing at least one field review of each field crew, working with the RCC to
ensure regional consistency, reviewing maps and field descriptions, and delivering the map products
to the Mapping Task Leader. The field crews, consisting of one or more soil scientists, mapped the
62
-------
Table 5-4. Population Estimates of Combined Number, Length, and Percentages of Stream
Reaches Having ANC Less Than Selected Reference Values at Downstream Nodes
During Spring Baseflow for NSS Phase I and DDRP M-APP Region
Sample (n)
Target Population
(N)
Length (km)
ANC < 0 u.eq/L
(N)
(%)°
Length (km)
(%)d
ANC < 50 u.eq/L
(N)
(%)
Length (km)
(%)
NSSa
164
25,715
69,569
326
1.3
2,324
3.3
2,199
8.6
7,430
10.7
DDRPb
36
5,496
15,239
181
3.3
597
3.9
1,504
27.4
3,002
19.7
a National Stream Survey estimates for Subregions 1D, 2Cn, and 2Bn combined.
b DDRP sample and target population for the Mid-Appalachian Region.
c Percentage of estimated target population number.
d Percentage of estimated target population length.
63
-------
watersheds, described the soils and soil map units, and transected randomly selected delineations of
map units to determine map unit composition.
Mapping protocols were developed in cooperation with soil scientists who worked in the
M-APP Region and were familiar with standards and procedures used in the National Cooperative Soil
Survey (Lammers et al., 1988). The mapping phase was scheduled for October 15, 1987, to July 30,
1988, in order to accomplish compilation and correlation tasks before the sampling phase, scheduled
to begin in September 1988. A preliminary regional soils identification legend was developed from
existing soil survey legends within major land resource areas in the region. The State Soil Scientist in
each state prepared a work plan to arrange for personnel and equipment to conduct the mapping.
Before mapping started, the Mapping Task Leader, Regional Coordinator/Correlator, State Office Soils
Staff, and Field Soil Scientists involved in the mapping participated in a training workshop to review
and practice the mapping protocols. The purpose of the workshop was to promote more consistent
interpretation and application of the protocols.
Map cartography and map compilation were conducted by SCS soil scientists. The area of each
map unit was estimated with a dot grid or pianimeter. The map symbol, map unit name, and area of
each map unit in acres was listed in a legend on each watershed map. These estimates of the area of
map units were used during correlation and selection of classes of soils for sampling. A more precise
measurement of the area of the map units was obtained when the maps were digitized from a rectified
map and entered into the GiS. Soil mapping activities and quality assurance of the mapping data
were described in depth in a report by Kern et al. (1990). \
5.4.1 Soils
Soils were mapped using standards and procedures specified in the National Soils Handbook
(Soil Survey Staff, 1983) and Soil Survey Manual (Soil Survey Staff, 1951). Soils were classified
according to Soil Taxonomy (Soil Survey Staff, 1975). Each map unit represented a collection of areas
defined and named the same in terms of their soil components, miscellaneous areas, or both, and
identified with a unique map symbol. Units that consisted of one dominant component (consociation)
and units with two or more dominant components (complexes) were mapped. Although most soil
components of map units were phases of soil series, some components were phases of soil families
or higher categories of taxonomic classes. The soil map units and soil components that make up the
map units were described with the following characteristics:
• name and symbol of the map unit
• regional landform
• local landform
• geomorphic position
• slope configuration
• percent composition of map unit components
• characteristics of the soil components
- name of the soil component
- phase (i.e., slope, texture, rock fragments)
64
-------
- drainage class
- permeability
- parent material, origin and mode of deposition
- depth to bedrock
- depth to impermeable layer
- taxonomic classification
Soil map units are cartographic delineations of the landscape that follow natural landscape
features and reflect the dominant soil conditions of the landscape segment. Map unit boundaries were
located by interpreting aerial photographs and verified by field observations. Soil mapping was based
on enough observations to determine soil/landscape relationships and confirm predictions of soil
occurrence established from these relationships. The soils in each map delineation were identified by
traversing and field investigation. Transects were conducted on randomly selected delineations of
map units to determine the proportionate extent of map unit components. The minimum size of an
individual map delineation was 3 ha, except for areas of hydric soils or soils formed in alluvium that
were delineated to a minimum of 1 ha. Inclusions may have been larger than the minimum if the soils
were similar and there were no readily observable landscape features to use for delineation. The
proportion of small areas of contrasting soil or miscellaneous area (inclusions in map unit delineations)
were estimated to the nearest 5 percent and were thereby included in aggregated values for a
watershed.
5.4.1.1 Soil Correlation
Soil correlation is the process of maintaining consistency in naming, classifying, and interpreting
soils and units delineated on maps. There are two main elements of soil correlation: (1) the correla-
tion of an individual soil pedon or groups of soil pedons with a soil series, or with some higher level
soil taxonomic class, and (2) the correlation of map units. Correlation requires consistent methods of
observation and measurement among all participants, as well as the use of consistent conventions
and terminology. The Soil Survey Manual (Soil Survey Staff, 1951) and Soil Taxonomy (Soil Survey
Staff, 1975) provide the conventions and guidelines for defining and naming map units, and for
defining diagnostic properties and taxonomic classes of soils used in the National Cooperative Soil
Survey. Soil series are defined by official soil series descriptions.
The soil correlation process started with the development of the preliminary regional identifica-
tion legend and continued throughout the mapping. The preliminary identification legend was based
on soil map units that had been mapped previously within the M-APP Region. These map units, there-
fore, had been tested for soil-landscape relationships and were expected to apply to much of the area
to be mapped. Consistent breaks for slope phases for map units and the use of the most common
soil texture phase for a soil series were established by the preliminary legend, in which 379 units were
listed. Soil map units were not limited to those in the preliminary legend and map units were
redefined and added as necessary during the progress of the field mapping. Including the map units
taken from the preliminary legend and those added during the mapping, 195 map units were used in
the mapping.
65
-------
5.4.1.1.1 Field correlation--
The soil scientist responsible for mapping each watershed performed the first level of correlation
of the soils and map units. The descriptions of official soil series were adapted to represent those soil
series for the region. At each point along a traverse, the soil scientist examined and evaluated the soil
for characteristics falling within (or similar to) the range of a soil series. Soils that were dissimilar to all
recognized soil series were classified at the family level in Soil Taxonomy (Soil Survey Staff, 1975).
Brief descriptions were made of the different kinds of soil, to document what was observed and to
compare or correlate with the official series description or other field descriptions. During the progress
field review, the State Soils Staff further evaluated the descriptions for consistent correlation within
each state.
In addition to the proper recognition and classification of soils observed, the soil scientist also
estimated the proportionate extent of the components that comprised each map unit. In this manner,
soil map units were described. The soil scientist then correlated the composition of soils with one of
the map units in the preliminary legend or proposed an additional map unit. The State Soils Staff
reviewed the map units to correlate map units on all watersheds within each state. The RCC con-
trolled the mapping legend and correlated soil map units throughout the M-APP Region.
5.4.1.1.2 Correlation and mapping review workshop -
A soil correlation workshop for the M-APP Region was held in Corvallis, Oregon, August 22-25,
1989, in conjunction with a workshop in which sampling classes were developed for the sampling
phase of the survey. The objective of the correlation segment of the workshop was to produce a
correct and consistent mapping database. This task included correlation of soil map unit components
and soil map units, resolution of map unit composition, completion of missing data, verification of soil
component and map unit data, and correction of map discrepancies. Correlation decisions made
during the workshop are documented by Kern et al. (1990).
5.4.1.2 Soils Database
Soil mapping of the M-APP Region generated vast amounts of data. In order to verify, validate,
and analyze these data, we entered them into computer database files. Data products generated by
the mapping included the identification legend, descriptions of the soil map units, descriptions of the
soil map unit components, soil transect information, and the soil map. This section describes the
database files developed for the DDRP mapping data and the procedures and QA/QC checks used
during development of the database. Data on map units in the preliminary legend and components of
those map units were entered into data files using a Mapping Data Management program with dBase
III plus software. Taxonomic classifications of soil series listed as components of map units in the
preliminary legend were added by merging the data from a magnetic tape of soil series of the United
States. The Mapping Task Leader at ERL-C had overall responsibility for data quality and for data
validation. The maps were digitized for input to a GIS at ERL-C as described in Section 5.4.7.
66
-------
5.4.1.2.1 Soil identification legend -
A preliminary soil identification legend was developed from existing soil survey legends within
the M-APP Region. Map symbols were assigned to map units in the legend by the RCC. Corrections
and additions to the legend were approved by a State Soil Scientist and the RCC and then added to
the identification legend (MA_MP_UN) file. Map unit symbols used in the mapping as shown on indi-
vidual watershed maps were entered into a preliminary watershed legend (MA_SHEDS) file prior to the
correlation workshop. Map units in the identification legend file that were not included in the
watershed legend file (not used in mapping) were marked for deletion. The identification legend was
listed for review during the correlation workshop and changes to the legend, determined during the
workshop, were made in the data file. Legends from each watershed map were also entered into the
GIS as the maps were digitized at ERL-C.
The legend data from the GIS were transferred to a dBase III file (MAGISOIL) where they were
summarized for the region and then compared to the regional soil identification legend. Discrepancies
were then resolved, and the map unit names were checked with the descriptions of the map units for
validity. The soil identification legend database file for the M-APP Region, named MA_MP_UN, con-
tained the following information for each map unit: map symbol; map unit name, including the name
of dominant soil components, texture modifier (e.g., gravelly, mucky), texture phase, slope phase, and
other phase {e.g., very stony, rocky); regional landform; local landform; geomorphic position; slope
shape across; slope shape down; and area in acres (determined from the GIS database).
5.4.1.2.2 Soil map unit composition and map unit components --
Map unit descriptions include both information on the proportionate extent of the different
components that make up the map unit and information on soil properties and classification of the
components. The same soil map unit component often occurs in more than one map unit. For
example, a major soil component in a consociation may have the same attributes as a major compo-
nent in a complex or a minor component in another map unit. Map unit data were separated into two
files, a map unit composition file and a soil components file, to eliminate duplication of data for the
same map unit component and facilitate retrieval of information. Each unique soil component was
assigned a component code to aid in accessing the soil-related properties and taxonomic classifica-
tion of a soil component. The map unit composition file named MA_MP_CM contains the map symbol,
soil name, component code for every component in the map unit, and proportionate extent of each of
the components.
The soil components file was named MACMPNT. Each record in this file included the compo-
nent code; soil name, texture, and slope of the component; five characteristics of the soil: permea-
bility, drainage, depth to bedrock, origin, and mode of deposition of the parent material; and the
taxonomic class. The sampling class code for the class with which the soil component was grouped
for sampling was added to the records in this database file following the sample class development
workshop. The records from the three database files, MA_MP_UN, MA_MP_CM, and MACMPNT, were
merged to display information about each map unit on a soil map unit worksheet. This worksheet
included the map symbol, map unit name, information about the landscape, major soil components,
67
-------
minor soil components, the proportion of each component in the map unit, and information about the
major components, including the taxonomic classification. Copies of these computer-generated
worksheets were reviewed by field soil scientists during the correlation workshop and corrected where
needed. Corrections made to the mapping database resulting from decisions made during the corre-
lation segment of the workshop have been documented by Kern et al. (1990).
5.4.1.2.3 Soil map unit transects -
Transect data were entered into files named MATRAN (for routine transects) and MATRANRC
(for transects conducted by the RCC). Each record in the transect files contained the watershed
name, watershed identification code, state responsible, map symbol, and transect number, along with
the soil name, slope gradient, and note for each of the 10 stops along the transect. Following the
workshop, a sampling class was assigned to the soil at each stop in the transect file by relating the
soil name to the same soil name in the components file and replacing the corresponding sampling
class code in the transect file with the sampling class code from the components file. In this manner,
misspelled soil name entries and soils not in the components file were identified by missing values in
the sampling class field of the transect file. These data were reviewed and corrected on the printout of
the transect file by soil scientists representing each state. Corrections to the computer data file were
made and any remaining discrepancies were resolved by the Mapping Task Leader and field soil
scientists.
5.4.2 Depth-to-Bedrock Mapping
Field soil scientists determined the depth to bedrock by examining and observing soil pedons,
readouts, excavations, and stream incisements. Maps of depth to bedrock within the watersheds were
drafted on mylar overlays of USGS quadrangle film positives at a scale of 1:24,000. The maps were
prepared by combining contiguous delineations of soil map units with the same depth-to-bedrock
class. One to three depth classes were assigned to each soil map delineation. The depth-to-bedrock
classes reflect major soils in the delineation, not map unit inclusions. The depth-to-bedrock symbols
for a map delineation are the depth-to-bedrock classes in descending order of estimated extent.
Where appropriate, soil map delineations were segmented into more than one depth-to-bedrock class.
Definitions of the depth-to-bedrock classes are in Kern et al. (1990).
5.4.3 Land Use and Forest Cover Mapping
Soil scientists prepared a land use/forest cover map of each watershed from field study and
photo interpretation. The mapping was transferred to a mylar overlay of USGS quadrangles at a scale
of 1:24,000. Land use was designated by the land use codes developed for the SCS soil description
form and as described by Kern et al. (1990). Forest cover types were those defined by the Society of
American Foresters (Eyre, 1980).
68
-------
5.4.4 Bedrock Geology Mapping
State bedrock geology maps were used to generate the bedrock geology coverages. These
were the only maps available that provided a consistent geologic classification scheme across the
region. The scale of the geology maps is 10-20 times smaller than the other mapping.
5.4.5 Drainage Mapping
Drainage features maps were drafted onto mylar overlays of USGS quadrangles at a scale of
1:24,000. Standard map symbols shown in the mapping protocols (Kern et al., 1990) were used to
show all perennial streams, dominant (well-defined) intermittent streams, water bodies, and other water
features. Marsh or swamp symbols were used to show areas of less than two acres with standing
water; larger marsh or swamp areas were delineated on the overlay and labelled "S". The upper ends
of perennial streams were marked with a short red line across stream symbols.
5.4.6 Quality Assurance
A rigorous QA/QC plan was implemented during all phases of mapping activities. QA/QC
activities included field reviews, independent evaluation of the mapping, and random transecting of
selected soil map units in the watersheds. The approach used in the transecting was described in the
mapping protocols (Lammers et al., 1988). Routine transect data collected by SCS field soil scientists
were used to make best estimates of the proportionate extent of map unit components. Transects
conducted by the RCC were used to evaluate the correctness of the soil map unit composition.
5.4.6.1 Field Reviews
Field reviews were conducted according to the guidelines of the NCSS (Soil Survey Staff, 1983).
Mapping activities of each crew were evaluated and the following items were checked: adherence to
protocols, base maps, map symbols, correspondence between soils and depth-to-bedrock mapping,
documentation of soil taxonomic units, soil map units, pedon descriptions, general notes, soil
transects, map delineation boundaries, and choice of map units. In addition, the reviewer conducted a
random traverse of the watershed to determine whether the mapping was correct.
5.4.6.1.1 Field reviews by the SCS state staff -
The SCS State Soils Staff and soil scientists in Pennsylvania and West Virginia conducted at
least one field review of each mapping crew as specified by mapping protocols. Results of the field
reviews are reported by Kern et al. (1990). No mapping was rejected by the reviewers and no major
problems were identified during the SCS field reviews. Most commonly discussed were the choice
between several similar map units, the size of delineations, changes to slope classes, the addition of
soil map units to the legend, and minor changes in vegetation mapping. Minor cartographic mistakes
were also found during the field reviews, which were useful to the field personnel for clarifying aspects
of the protocols.
_
69
-------
5.4.6.1.2 Field reviews by the RCC -
As specified in the mapping protocols (Lammers et al., 1988), the RCC participated in one field
review of a watershed for each State Soils Staffs region of responsibility. The RCC conducted one
field review in Pennsylvania and one in West Virginia in conjunction with reviews by the SCS State
Soils Staff personnel. These reviews are discussed in more detail in the final report of the RCC,
included as Appendix D in the report by Kern et al. (1990).
5.4.6.2 Evaluation of Mapping by the RCC
Eight watersheds were randomly selected for evaluation on the basis of one watershed for each
soil survey crew. Within the eight watersheds, map units were selected for transecting by the RCC.
Mapping was evaluated by examination of stereoscopic pairs of aerial photographs. Relationships
between soils and landforms were scrutinized and areas were identified that required closer examina-
tion by traversing and transecting. The RCC examined the mapping in delineations that were selected
for transecting and areas that were traversed while going to or coming from the transects. The goal
was to evaluate 20 - 50 percent of the map delineations in the selected watersheds. Map correctness
was Judged by how well the map unit descriptions fit the soils in the mapped areas, and how accu-
rately map unit boundaries were drafted on the maps. Watershed boundary location, soil mapping,
land use, and forest cover mapping were evaluated. The depth-to-bedrock mapping was not evalu-
ated separately because it was based on the depths of soil map unit components. Mapping was
judged as acceptable as mapped on seven of the eight watersheds. The RCC requested that the soil
map for one watershed be revised to include the delineation of small areas of alluvial soils that were
greater than one hectare in size. The soil map for this watershed was judged as acceptable after
revision. Kern et al. (1990) have presented the report by the RCC summarizing his activities and
findings.
5.4.6.3 Evaluation of Soil Transect Data
Transects consisting of 10 observations of the soil at equally spaced intervals across a map unit
delineation were conducted by field soil scientists during the mapping and by the RCC during evalua-
tion of the mapping. The transects conducted by the field soil scientists, referred to as routine
transects, were used in estimating the relative composition of map unit components in a map unit.
The design for selecting routine transects resulted in one transect being conducted on most of the
map units in a watershed and is discussed in Lammers et al. (1988). This approach resulted in a total
of 351 routine transects (3,510 data points), of which 348 were used for analysis. Of the 195 map
units in the survey area, 37 (19 percent) had 3 transects or more, 40 (20 percent) had 2 transects, 92
(48 percent) had 1 transect and 26 (13 percent) were not transected.
Transects conducted by the RCC were used to evaluate the correctness of map unit composi-
tion that was estimated during the mapping and correlation based on field observation and transect
data. The RCC conducted transects in the watersheds that were selected for mapping evaluation.
Map units selected for transecting by the RCC had three or more routine transects. The RCC
conducted 61 transects in 24 different map units. The routine transects do not provide an indepen-
70
-------
dent evaluation of the accuracy of the proportions of soil components, because the information from
the routine transects was used in the soil correlation workshop to adjust the proportions of map unit
components in the map unit composition database. The routine transects, however, do provide a
means of evaluating whether the delineations within the same map symbol are alike in a proportion of
components.
The proportions of soils transected by the RCC were compared to the expected proportions in
the map unit composition database to determine if there were any significant differences. Also, routine
transects were compared to RCC transects. Finally, soil components observed at transect points were
assigned to the appropriate sampling class, and map unit correctness was evaluated with respect to
the sampling class composition. The latter evaluation was especially relevant in judging the
correctness of map units for the purposes of the DDRP. All statistical tests were conducted using two-
sided alternative hypotheses.
5.4.6.3.1 Analysis of the map units by major components -
Major components, those comprising 20 percent or more of the map unit, were determined from
the map unit composition database. For the major components of each map unit, the transect points
were treated as observations from a binomial distribution with the population proportion, p, where p
was calculated from the map unit composition database. The binomial distribution was used because
the outcome of each transect can be classified in exactly one of two ways; either the proportion of a
particular major map unit component matches the expected result from the map unit composition
database or it does not. The sample proportion used was the total number of transect points with the
same name as a major component, divided by the total number of transect points.
For the 348 routine transects used in the analysis, hypothesis tests were performed at the .05
level of significance to locate transects that had proportions of major components significantly different
from the expected proportions. There were 18 routine transects with proportions significantly different
from the expected. Under the null hypothesis of no discrepancies, we would expect 5 percent, or
roughly 17 transects, to be different.
After excluding significant individual transects, we calculated the variability in proportion of major
components between transects to determine if some soils not distinguished by the previous analysis
were highly variable in their proportions of major components. We calculated the variance of these
proportions for each map unit with more than one transect. Because the distribution of these vari-
ances was asymmetric, a robust data analysis technique was used.
We used a boxplot of the variances of proportions to assist in our analysis. Boxplots use the
interquartile range (IQR). This is the distance between the first and third quartiles, that is, the 25th and
75th percentiles, respectively. Points more than 1.5 times the IQR away from the closest quartile are
considered outliers, and points more than 3 times the IQR away from the closest quartile are consid-
ered strong outliers. No map units had variances of proportions that would be considered outliers or
strong outliers. This suggests that most of the potential problems had already been detected by
analyzing the individual transects.
71
-------
Similarly, we examined the RCC transects for those with significantly different proportions at the
0.05 level of significance. Six transects out of 62 had significantly different proportions. Under the null
hypothesis, we would expect an average of about three transects. Three of the transects had
extremely low proportions because the RCC used a different series name than that in the map unit
composition data file.
We also examined the RCC transects for proportions of major components in map units across
all RCC transects. The significant individual transects were removed first, using 0.05 as the level of
significance. Only two map units were significantly different. The number of transects expected under
the null hypothesis of no discrepancies would be approximately one or two.
5.4.6.3.2 Comparison of the routine and RCC transects -
For these comparisons, the standard two-sample test for a difference in proportions using the
normal approximation was used whenever the previously mentioned criteria were met for both
samples. When either sample was not appropriate for a normal approximation, we used Fisher's exact
test (Fisher, 1935).
One map unit had a significantly different proportion from that in the database, but did not have
significantly different transects for either the routine or the RCC transects. The two proportions appear
as significantly different because the routine proportion is slightly above the proportion calculated from
the RCC transects, whereas the RCC transect proportion is slightly below the expected proportion.
The expected number of significant map units under the null hypothesis is also one. This strong
agreement suggests that the routine mapping produced results that were not observably different from
those of the RCC.
5.4.6.3.3 Analysis of the map units by sampling class •-
In the DDRP, soils thought to be similar with respect to acidic deposition were grouped into
classes (see Section 5.5.1). Because there were several hypothesis tests for each map unit, we used
the Bonferroni inequality (Johnson and Wichern, 1982) to handle the error rate of the simultaneous
hypothesis tests within each map unit. This was a conservative approach, and the simplest to
program and implement.
After removing significant individual transects, we analyzed the RCC transects for significant
sampling classes within map units. Three different map units had sampling classes that differed sig-
nificantly from the proportions estimated by the soil scientists. At the 0.05 significance level, we would
actually expect one or two map units with significant sampling classes under the null hypothesis. The
three map units observed is not an unusual occurrence, inasmuch as the probability under the null
hypothesis of observing three or more map units is approximately 13 percent.
After removing significantly different individual transects, the routine transects were also analyzed
for significant sampling classes within map units. Out of 170 map units analyzed, only 8 had signifi-
cant sampling class proportions. Under the null hypothesis, about 8 or 9 map units would be
72
-------
expected. This agreement suggests that for the most part, the sampling classes for the map units are
not noticeably different from their expected proportions.
5.4.6.3.4 Conclusions of the transect analyses --
The transect analyses suggest that, after removing significant transects, no map units had a
composition of unusually high variability. After the significant transects were removed, the map units
were very consistent. The sampling classes matched the expected proportions well, indicating that
there were few problems in the mapping as far as use by the DDRP was concerned.
5.4.6.4 Exit Meeting
An exit meeting was held to review and document mapping operations. Individual differences in
interpretation of protocols and the need to make on-the-spot decisions during field operations has
been recognized by project scientists. These activities may affect the quality of the mapping or the
interpretations applied to the mapping data. A review of mapping operations pointed out some weak-
nesses in mapping protocols, training, and communication and-helped to evaluate mapping quality
and reliability of assessment results. The review was conducted by soliciting responses to a
questionnaire from SCS soil scientists representing the states responsible for the mapping, the
Mapping Task Leader, and the RCC.
5.4.7 Geographic Information Systems Data Entry
The DDRP obtained data from contract mappers and from existing information. The USDA-SCS,
in cooperation with the U.S. EPA, mapped soils, vegetation/land use, drainage, and depth to bedrock
at a scale of 1:24,000 for each watershed (Lee et al., 1989a) (see Sections 5.4.1 through 5.4.5).
Bedrock geology was extracted from existing state geology maps. These data were all entered into a
GIS using ARC/INFO software. The GIS is designed to automate, manipulate, analyze, and display
geographical data in digital form, and was used in the DDRP as a spatial tool for technical analysis
and for effective communication (Campbell et al., 1989). Church et al. (1989) presented examples of
GIS maps of watershed characteristics for a specific watershed.
Much of the M-APP mapping database was developed in the same way as the NE and SBRP
databases (Church et al., 1989). The M-APP database followed the same QC procedures to ensure
the data were entered as consistently and accurately as the original mapped information allowed.
These procedures included checking the accuracy of the information before, during, and after
digitizing. Two independent checks were performed at critical stages of the digitization before the
data were accepted for use within the DDRP. Mortenson (1989a,b) has provided extensive details
concerning the entry and checking of mapping data in the GIS for the DDRP.
5.5 SOIL SAMPLING PROCEDURES AND DATABASES
Soils were described and sampled to provide the morphological, physical, and chemical data
needed for the DDRP analyses. In the M-APP, 150 pedons were described and 900 samples (i.e.,
73
-------
about six horizons per pedon) were taken. A pedon is the smallest volume of soil that has all the
characteristics by which a specific soil is defined. Operationally, it is usually taken to be about a meter
square in cross section to depth of 2.0 m or to bedrock, whichever is shallower. Soil survey activities
have been described in some detail by Lee et al. (1989a).
5.5.1 Development/Description of Sampling Classes
5.5.1.1 Rationale/Need for Sampling Classes
As in the NE and SBRP, many more soils were mapped in the M-APP than could practically be
sampled. The approach used by DDRP in each region was to map and correlate soils in each DDRP
watershed to a single regional legend and then group the soil map unit components (hereafter called
"mapped soils") expected to have similar chemical and physical characteristics with respect to their
responses to acidic deposition into regional sampling classes. The criteria used to group the mapped
soils were formalized into a key, which then provided the basic definition of the classes, superseding
the specific taxa used to develop it. The intent was for the key to be applicable to every soil in the
target population (Church, 1989) of watersheds in the region.
Several pedons in each sampling class were described and sampled to determine the means
and variances of the characteristics of each sampling class. A randomized design was used to select
pedons for sampling. Within the context of the DDRP in the M-APP Region, all soils in a given class
were considered to be similar. Sampling crews judged a soil to be correct or incorrect for sampling
on the basis of the sampling class into which it was grouped, not by which series or map unit compo-
nent (if any) it represented. The regional means and variances were then used in conjunction with the
soil maps to build area or volume-weighted estimates, with error estimates, of each watershed's soil
characteristics. These estimates were then used in DDRP analyses. Results from the NE and the
SBRP have been presented by Church et al. (1989). Details of DDRP watershed mapping and soil
sampling were presented by Kern et al. (1990) and by Kern and Lee (1990), respectively. The role of
sampling classes in DDRP was described by Lee et al. (1989a).
In the M-APP Region, available resources limited soil sampling and laboratory analysis to 150
pedons. This implied an average sampling density of 4.2 pedons per watershed, or 1 pedon per
190 ha. Thus, the purpose of the sampling classes was to attain adequate resolution of soil
properties, while limiting the sampling effort to a practical number of pedons.
5.5.1.2 Approach Used for Sampling Class Development
Sampling classes have previously been defined for the NE (Lee et al., 1989b) and the SBRP.
Sampling classes were developed for the M-APP Region at a workshop held after soils, vegetation,
land use, and drainage pattern had been mapped (1:24,000) on the DDRP watersheds by SCS soil
scientists, and after data on the occurrence and composition of soil map units had been entered into a
database (Lammers et al., 1987; Lee et al., 1990). Participants included the SCS soil scientists from
Pennsylvania and West Virginia who had the greatest responsibility for DDRP M-APP mapping and
74
-------
subsequent sampling; a representative of the SCS National Quality Assurance staff; DDRP scientists
and statisticians; and other scientists with related interests.
5.5.1.3 Description of Mid-Appalachian Sampling Classes
The M-APP sampling classes are defined by Figure 5-3. Except for "Mine Soils" and drainage
classes, the terms used in Figure 5-3 are as defined in Soil Taxonomy (Soil Survey Staff, 1975). This
includes "fragipan" ("Fragi" Suborders) and "Flooded" ("Fluv" root in Subgroup name).
The Entisols have the weakest development among the M-APP soils; specifically, they do not
have diagnostic B horizons. Frequently, they do not have B horizons of any kind. This is important for
DDRP, because in some applications, aggregation by major horizon is the first step in aggregation of
the soil data to the watershed level (Church et al, 1989). Also, separating Entisols allows greater
precision for soils that might have effects disproportionate to the area they occupy. For example,
soils formed from mine spoils are likely to be very heterogeneous and, possibly, very acidic; DDRP
has at least some data for these soils for every watershed in which they occur. The flooded soils
(Fluvents, Fluventic Dystrochrepts, Fluvaquentic Dystrochrepts) are important for streamwater chemis-
try because of their proximity to the streams. They are likely to differ from other soils because of the
alluvial deposition/removal processes. Differences might include surface disturbances, irregular
decreases in organic carbon with depth, and/or relatively high organic carbon at depth.
The organic parent material of the Histosols makes them distinctly different from the mineral
soils in physical and chemical properties. Histosols also tend to be located near surface waters, and
thus might exert an influence on surface water chemistry disproportionate to the area they occupy.
Defining a Histosol sampling class (HST) ensured that, despite their small extent, the DDRP has data
on these potentially important soils.
The Ultisols and Alfisols both have an argillic horizon, but are distinguished by Ultisols having
less than 35 percent base saturation (Soil Survey Staff, 1975). Because Alfisols in this region might not
have much more than 35 percent base saturation, they were grouped with the Ultisols. The clay-
enriched argillic horizon of these orders is likely to have relatively low hydraulic conductivity, thus
influencing the flow of water to streams. Also, these horizons are likely to differ from nonargillic
horizons in properties correlated with clay content, such as cation exchange capacity (CEC) and
sulfate adsorption.
The presence of a fragipan restricts water movement, thus limiting the amount of soil available
for interaction with precipitation, and, possibly, the time the precipitation is in contact with soil as it
flows to the stream. Soils that are shallow or skeletal have similar inherent limitations. Compared to
other soils, these soils might be limited in their ability to neutralize acidity from precipitation.
Soils assigned to the BMK class (skeletal, moderately deep Typic Dystrochrepts) occur on
approximately 8,900 ha, or almost one-third of the total area mapped in the region. The workshop
members discussed the possibility of splitting this class into less extensive classes, but could not
identify any meaningful soil characteristic to use.
75
-------
FLOC
SOME!
POO
DRAI
&WE
ENTISOLS
Ml
SO
>DED
Few
WHAT WEI
RLY MOD.
NEDR DRA
CFP FL
1
Ix
! SOILS OF THE MID-APPALACHIANS ||
' XX
j^sBHssss^sBHHHHs^s^HBHs*.^*54***
1
HISTOJ
[H
•JE
LS
71 FLOC
»**J
JL&
WELL
NED
_J
SOLS
_
El r
' NO
FRAGIP
H
>DED NC
FLOO
1
SHALLOW MC
DE
| BNS |
1
SKELETAL
I
KXXXXXXXXXXXXXXXXXXXXXXXXXXXX:
1
INCEPTISOLS
AND
SPODOSOLS
1
i r
N(
AN FRAGIPAN FRAG
1
1 1
OCHREPTS AQUEPTS
DED 1 BXWJ | BXP |
1
)D. DEEP
bH 1
WELL &
1 MOD. WELL
| END | DRAINED
ULTISOLS
AND
ALFISOLS
"PAN FRAGIPAN
1
SOMEWHAT WELL &
POORLY MOD WELL
DRAINED DRAINED
| TXP | | TXW |
1
SOMEWHAT
POORLY
DRAINED
& WETTER
MOD. _,_. 1
nccp DEEP i — < 1
NON i TPD
SKELETAL 1 ' '
BMK
TWM
TWO
Figure 5-3. Definition of soil sampling classes for the DDRP Soil Survey in the M-APP Region.
76
-------
Soils that are "somewhat poorly drained and wetter" almost always have an aquic moisture
regime. They have water at or near the surface for prolonged periods of time, and are likely to have an
important role in the flow of water to the stream. Aquic soils are chemically different from other soils in
that they tend to have reducing environments for prolonged perjods. Sulfate entering these soils could
be reduced to sulfide, thus reducing the potential for acidification of surface waters. Conversely, drying
of these soils (e.g. through artificial drainage) might release sulfate through oxidation of sulfide,
causing an increased rate of acidification.
An indication that the sampling classes succeed in being applicable to all soils on the DDRP
watersheds comes from the transect observations made as part of mapping quality assurance (Kern et
al., 1990). About 3500 point observations were made prior to the development of the sampling
classes. All the soils described at these points were subsequently assigned to sampling classes,
including soils that had not been identified as map unit components. The extent of soils on the DDRP
watersheds that do not fit any sampling class must therefore be quite small.
5.5.1.4 Comparison of Sampling Class Schemes for the Three Regions
The distinctions among the sampling classes defined for the M-APP Region (Figure 5-3) reflect
differences in soil properties expected to be important for the effect of acidic deposition on streams.
Base cation supply, sulfate adsorption, water flow, and proximity to streams were important considera-
tions. These are the same items that were considered in defining DDRP sampling classes for the NE
(Lee et al., 1989b) and SBRP. Thus, the sampling schemes for-the three regions are consistent in
rationale and purpose. The differences among them resulted from tailoring the schemes to fit the
specific soils that were mapped on DDRP watersheds in each region.
Origin of parent material was an important differentiating criterion of the sampling classes for the
NE (Lee et al,, 1989b) and SBRP. It was not used in the M-APP Region because, except for two
watersheds in the Piedmont and Blue Ridge Mountains of Virginia, all DDRP M-APP watersheds are on
sedimentary bedrock. A separate sampling strategy was developed for the two nonsedimentary
watersheds (Lee et al., 1990, Section 5.5.2).
The use of drainage class to segregate sampling classes was nearly equivalent to using taxo-
nomic criteria (i.e., moisture regimes). Entisols could have been separated by "Aquents" versus
"Fluvents" without altering the mapped soils assigned to classes CFP or FLW. Similarly, Ultisols/Alfisols
without fragipans could have been separated by "Aquults/Aqualfs" versus "Udults/Udalfs" without
altering the mapped soils assigned to classes TXP or TXW. Except for one mapped soil (Wharton,
Aquic Hapludults), the Ultisols/Alfisols without fragipans could have been separated by "Aquic
Suborder or Subgroup" versus "Other." Thus, workshop participants relied mainly on the taxonomic
concept of aquic versus nonaquic suborders or subgroups to differentiate wetness, and used drainage
class mainly as an expedient method for sorting soils into groups by computer. This was consistent
with the approach used in the NE (Lee et al., 1989b). In that region, taxonomic criteria were preferred
to drainage classes because they were considered to be better defined, more objective, and more
consistently applied in the field. Neither drainage nor moisture regime was used to define sampling
classes in the SBRP.
77
-------
The scheme for the sampling classes in the M-APP is similar to that for the NE (Lee et al.,
1989b) in that it relies heavily on the hierarchy and formative elements of Soil Taxonomy (Soil Survey
Staff, 1975). For example, both schemes have Order as the initial criterion. In comparison, the
scheme for the SBRP relied more on soil-related properties. Mode of deposition of parent material
was explicitly included as a differentiating criterion in the NE (i.e., glaciofluvial, glacial till, alluvial), and
was implicit in the "flooded" (i.e. alluvial) classes in the M-APP and SBRP schemes. Soil depth and
family particle sfee distribution were used as differentiating criteria in all three DDRP regions. The
presence of a slowly permeable layer was used to differentiate classes in the NE (permeability class)
and in the M-APP (fragipans), but was not used in the SBRP.
The use of "Ultisols/Alfisols" as a proxy for the presence of an argillic horizon in the M-APP was
possible only because no other soils with argillic horizons were mapped. Thus, it is an example of a
criterion that is specific to a region. Similarly, "Spodosols" was an important criterion in the NE,
whereas Spodosols and Inceptisols were grouped together in the M-APP (classes BMK and BND) and
in the SBRP (class FR, "Frigid Soils"). The use of these criteria emphasizes that the sampling classes
are not universal, but apply only to DDRP watersheds within the respective regions.
5.5.2 Selection of Sampling Sites
There is a strong tendency for soil scientists to select typical soils for sampling. Although this is
proper for most applications, it would not have been appropriate for the statistically based sampling
scheme used by the DDRP. To ensure an unbiased sample for estimating means and variances of the
characteristics of sampling classes over the regions, the DDRP used a unique three-part scheme of
randomly selecting sampling sites for each sampling class: (1) selection of watersheds from those in
which the desired sampling class occurred, (2) random location of potential sampling sites on soils
maps within delineations in which the class occurred (Figure 5-4), and (3) random selection of transect
direction if the field crew found that the desired sampling class did not occur within 5 m of the
potential sampling site (Figure 5-5).
Most of the sampling classes are broadly distributed across the region. The major exceptions to
this are those classes dominated by soils developed in glacial parent material (classes BXP and BXW);
these soils are limited to the extreme northern part of the region. Thus, the properties (e.g., sulfate
concentrations) of soils within individual classes may vary along the deposition gradient in the region.
This possibility was not incorporated into the sampling classes, but was accommodated in the alloca-
tion of sampling effort. Five clusters of watersheds were identified based on geologic, climatic, and
deposition differences. Cluster 1 contained watersheds in the glaciated part of northeastern
Pennsylvania [Glaciated Allegheny Plateau and Catskill Mountains Major Land Resource Area (MLRA)]
and the northernmost part of the Northern Appalachian Ridge and Valley MLRA (i.e., in subregion 1D
of the National Stream Survey; Kaufmann et al., 1988). Cluster 2 contained watersheds in the
remainder of the Northern Appalachian Ridge and Valley MLRA. Cluster 3 contained watersheds in
the Northern and Southern Appalachian Ridge and Valley MLRAs; one watershed was transitional to
the Blue Ridge MLRA. Cluster 4 contained watersheds in the northern portion of the Eastern
Allegheny Plateau and Mountain MLRA. Cluster 5 contained watersheds in the southern portion of the
Eastern Allegheny Plateau and Mountain MLRA, and one watershed in the Southern Appalachian
78
-------
Enter
J — A Desi
Designate sampling class!
Obtain soil map for
watershed selected for sampling
Designate map unit delineations
with at least 20% sampling class
Randomly position dot grid
on soil map
Number all points falling
on designated delineations
Determine percent composition (X)
of sampling class in map unit
Randomly select an integer (Y)
where 1
-------
c
Enter
i Obtain watershed map showing",? /
^ pre-selected starting points f/
M and vegetation classes |'J
YES
}
Proceed to point 1
Examine area within
5 meters of point
Desired soil and
vegetation found?
Randomly select
direction (N, NE, E •••)
i
r
Proceed 10 meters In
selected direction
"\
-^
Proce
F
\
point
150 meters
from starting
point?
Desired soil
and vegetation
found within 5 meters
of new point?
5 directions
tried?
All 5
pre-selected
oints tried?
Describe and
sample soil
Cannot sample class
on watershed
Figure 5-5. Field selection of a sampling point for sampling class on a watershed.
80
-------
Ridge and Valley MLRA. Table 5-5 shows the watersheds by cluster. The clusters were used only for
the widely occurring sampling classes END, BMK, and TXW. For almost all other sampling classes, at
least one pedon was sampled in all watersheds with at least 10 ha of soils in the sampling class. The
sole exception was 1 watershed with 14 ha of soils in class TWM. (See Lee et al., 1990, for details of
allocation of sampling effort to sampling classes, geographic groups, and watersheds.)
After a watershed was selected for sampling of a particular class, potential sampling sites
(usually five) were determined by random selection from grid points that fell on delineations of map
units with at least 20 percent of their area occupied by soils within the class (Figure 5-4). The
vegetation map unit (i.e., Society of American Foresters [SAF] cover type; Eyre, 1980) at each selected
point was noted and classified into one of five broad groups (conifer, hardwood, mixed, open dryland,
open wetland). A detailed description of the selection of potential sampling sites was documented by
Lee et al. (1989a).
In each watershed selected for sampling, the field crews proceeded to the first of the potential
sampling sites and determined whether a soil within the desired class occurred within 5 m (Figure
5-5). If there was any such soil, and if the vegetation at the site "fell into the broad group identified
from the vegetation map, the soil was sampled. Otherwise, the crew leader used a random number
table to select a transect direction. The crew proceeded in this direction, stopping at regular intervals
to determine if a suitable soil was present. They sampled the soil at the first site they found that met
the criteria for the sampling class and for the broad vegetation class. If no such site was found on the
first transect, another direction was selected (see Figure 5-5). If the desired combination of soil
sampling class and broad vegetation group was not found after five transects, the crew proceeded to
the-second preselected potential sampling point, until all preselected points on the watershed were
exhausted. In the vast majority of cases, sampling crews were able to locate a suitable soil at the first
potential sampling site. The instructions given to the crews for selecting sampling sites were
documented by Kern and Lee (1990).
A different approach was used for the two Piedmont/Blue Ridge watersheds. Sampling sites
were selected to be representative of important soils on those specific watersheds. ERL-C scientists
located sampling sites on the soil maps based on hydrologic relation to the stream, extent of soils in
the watershed, and distribution of sampling sites across the watershed, and specified the soil series to
be sampled at each site. Field crews went to each site and sampled a soil that they considered to be
representative of the series in that portion of the watershed.
Four pedons were sampled on each of the two watersheds. Subsequent evaluation of the
laboratory data for these eight pedons confirmed the expectation of distinct soil chemistry on the two
watersheds. Base cations and pH for most of these pedons were markedly higher than for the M-APP
sampling classes to which they would have been assigned. Chemistry data for four of the pedons
(Myersville, Eubanks, Eubanks Taxadjunct, Lew Taxadjunct) were similar, and were grouped for DDRP
purposes for these two watersheds. The other four pedons (Codurus Taxadjunct, Ashe, Cataska,
Catoctin) were dissimilar, and were not grouped. Data for these five soil units (i.e., the group of four
pedons and the four ungrouped pedons) were then used in the same way sampling class data were
used on the other watersheds.
81
-------
Table 5-5. List of DDRP Watersheds by Geographic Group
Group 1: 1D029023, 1D029031, 1D029042, 1D029043, 1D036011, 1D036017, 1D032005
Group 2: 2B036028, 2B036046, 2B036062, 2B041008
Group 3: 2B047032, 2B047036, 2B047076
Group 4: 2C028069, 2C028070, 2C028075, 2C029002, 2C029016, 2C029020, 2C035027,
2C041002
Group 5: 2C041039, 2C041040, 2C041045, 2C041051, 2C046005, 2C046033, 2C046034,
2C046041, 2C046050, 2C047007, 2C047010, 2C057004
Blue Ridge/Piedmont: 2B047066, 2B047089.
82
-------
_
5.5.3 Soil Sampling
The USDA Soil Conservation Service conducted the soil sampling activities for the DDRP. State
offices involved were Pennsylvania and West Virginia.
5.5.3.1 Soil Sampling Procedures
Protocols for DDRP soil sampling were developed for the M-APP (Kern and Lee, 1990) by
adapting the procedures of the National Cooperative Soil Survey (Soil Survey Staff 1975, 1983, 1984).
To enhance regional consistency, standard supplies and equipment were provided for the field crews
by the Soil Preparation Laboratory established in Las Vegas, Nevada. The crews shipped soil samples
to this laboratory, and obtained new supplies from it as needed. Laboratory personnel inspected the
samples for obvious problems (e.g., inadequate sample volume, poor labeling, possible contamina-
tion), thereby providing an additional check on regional consistency.
The protocols gave detailed instructions on the randomized procedure for locating sampling
sites (see Figure 5-5), for excavating pedons in difficult situations, and for documenting the site and
pedqn with notes and photographs. Soil profile descriptions were entered onto a form (SCS 232)
designed to facilitate entry into a database. Field estimates of percent rock fragments, included in the
profile descriptions, were used to correct for nonsoil volume during data aggregation (see Sections
8.8.3, 9.2, 9.3 of Church et al., 1989). Crews sampled every horizon thicker than 3 cm thick down to
bedrock or 2.0 m. Thick horizons were split for sampling. Samples were cooled to 4°C within 12
hours, and then shipped to the preparation laboratory via express courier.
An important change in sampling protocols for the M-APP from the NE and SBRP was the use of
alternative techniques for obtaining bulk density samples. The DDRP used clods as the primary bulk
density method because this commonly used method is relatively easy to use in the field. It was the
only method used in the NE and SBRP, where completeness was only 48 percent and 61 percent.
Two alternative methods (volume-replacement and volume-fill) were developed for the M-APP to
attempt to achieve 100 percent completeness. For volume-replacement, a roughly circular hole
(approximately 10 cm diameter and 10 cm depth) was excavated and filled with fiber reinforced epoxy
beads to determine its volume. For volume-fill, soil was excavated into a container of known volume.
For both methods, bulk density was calculated by dividing the oven dried weight of the fine soil by the
volume of the excavation, corrected for rock fragments. For the purposes of comparison, some hori-
zons were sampled using both clod and volume-fill or clod and volume-replacement methods. The
alternative methods were described by Kern and Lee (1990). Of these procedures the clod method
gives slightly higher values (J. Kern, pers. comm.). Completeness of bulk density samples was almost
100 percent for the M-APP (88 percent clods, 11 percent volume-fill, < 1 percent volume-replacement).
5.5.3.2 Quality Assurance/Quality Control of Sampling
The purpose of the QA/QC tasks for sampling was to ensure and document that the samples
were collected, handled, shipped, and documented in a consistent, proper manner. The QA/QC
procedures for sampling were described and evaluated by Kern and Lee (1990).
83
-------
Crews were trained at a regional workshop prior to sampling. During sampling, every crew was
audited by the State Soils Staff and the Regional Correlator/Coordinator, who were responsible for
consistency within each state and within each region, respectively. At least one site per state was
audited jointly by the State Soils Staff and the RCC.
Each crew also was audited by a member of the DDRP QA staff. As an independent evaluation,
the EPA auditor used a detailed checklist to document adherence or deviation from protocols as given
in the DDRP sampling manuals. Regional consistency was also promoted by feedback from the
preparation laboratory.
The QC activities also provided unique information on the variability of pedon descriptions
prepared by different soil scientists. The State Soils Staff and the RCC each performed independent
descriptions of pedons that also had been described by the sampling crews. Thus, for some pedons,
up to three independent descriptions were available. The primary purpose was not to decide which
soil scientist was "right", but to document the variability inherent in a procedure that is somewhat
subjective. Comparison of descriptions also was useful to promote consistent application of soils
concepts within states and regions. As an additional QA/QC check, the pedon descriptions were
reviewed for consistency by the SCS state offices and by EMSL-LV staff. Discrepancies were docu-
mented and resolved by consulting the field crews.
For every third pedon, each crew sampled one horizon in duplicate by placing alternate
trowelfuls of soil into two sampling bags. The same day that the crews collected duplicate samples,
they also prepared audit samples. These were 500 g soil samples supplied by the preparation labora-
tory that were carried into the field, sieved, and bagged as though they were routine samples. The
variability of the duplicate and audit samples was documented by Byers et al (1990).
5.5.4 Physical and Chemical Analyses
The chemical and physical analyses performed on DDRP soil samples are summarized in Table
5-6.
5.5.4.1 Preparation Laboratories
A preparation laboratory was established at the U.S. EPA Environmental Monitoring Systems
Laboratory in Las Vegas (EMSL-LV) to facilitate processing of the field-moist bulk soil samples col-
lected by the sampling crews and to perform preliminary analyses on these samples. Responsibilities
of this laboratory included air-drying, disaggregating, and sieving soil samples; measuring field-moist
pH, organic matter (loss on ignition), air-dry moisture, and rock fragments; and homogenizing and
subsampling the samples to produce analytical samples.
Analytical samples were grouped into batches of approximately 40 samples. Measurement qual-
ity samples, including field duplicates, field audits, natural audit samples, and preparation duplicates,
were placed in each batch for QA purposes. The measurement quality samples were packaged and
labeled in the same way as routine samples, and were not identifiable by the analytical laboratories.
84
-------
Table 5-6. Laboratory Analysis of DDRP Soil Samples
Chemical Analyses
1. pH (distilled water; 0.01 M CaCI2 ; 0.002 M CaCI2)
2. Total carbon
3. Total nitrogen
4. Total sulfur
5. Cation exchange capacity
a. 1 N NH4 OAc, pH = 7.0
b. 1 N NH4 Cl, unbuffered
6. Exchangeable bases (Na, K, Mg, Ca)
a. extraction by 1 N NH4 OAc, pH = 7.0
b. extraction by 1 N NH4 Cl, unbuffered
c. extraction by 0.002 M CaCI2
7. Exchangeable acidity
a. BaCI2 -TEA method, pH = 8.2
8. Extractable iron and aluminum
a. sodium pyrophosphate
b. ammonium oxalate
c. citrate-dithionite
d. 0.0002 M CaCI2
e. Aluminum by 1M ammonium chloride
9. Extractable sulfate
a. water soluble
b. phosphate extractable
10. Sulfate adsorption isotherms (six points)
11. Extractable silicon by ammonium oxalate-oxalic acid
Physical Analyses
1. Particle size (5 sand fractions, 2 silt fractions, clay)
2. Bulk density
3. Moisture content
85
-------
Organic and mineral soil samples were placed in separate batches, based on loss-on-ignition. The
preparation laboratory also measured the mass of all bulk density samples and the volume of the clod
samples, and calculated bulk density. Details of the preparation laboratory procedures were given by
Bartling et al. (1988).
Preparation laboratories were audited by DDRP QA staff before becoming operational and again
while operational. The QA/QC procedures for the preparation laboratories were described by Papp et
al. (1989). QA/QC results were evaluated by Papp and Van Remortel (1990), who concluded that soil
sample integrity was maintained at the preparation laboratories.
5.5.4.2 Analytical Laboratories
5.5.4.2.1 Analyses --
The analytical laboratories were contracted to perform the physical and chemical analyses listed
in Table 5-6 and described in Table 5-7. More complete descriptions of the procedures used by the
analytical laboratories were given by USEPA (1988).
In addition to the parameters listed in Table 5-7, a number of calculated variables were derived
for use in various analyses. Derivation of these variables is described in sections where the variables
are first used (e.g., sulfate adsorption isotherms in Section 9.2 and cation exchange selectivity
coefficients in Section 9.3).
5.5.4.2.2 Selection of analytical laboratories -
The solicitation process (Byers et al., 1990) began with preparation of detailed statements of
work (one for total N, C, and S; one for all other analyses) that defined the analytical and QA/QC
requirements in contractual format, followed by preparation and advertisement of an invitation for bid.
All interested laboratories were sent performance evaluation (PE) soil samples to be analyzed
according to DDRP procedures; these samples had been previously characterized for DDRP. Bidding
laboratories were rated using a scoring sheet developed by Papp et al. (1989). For analyses other
than total C, N, and S, all laboratories that passed the PE sample evaluation were then audited to
verify their ability to meet the contractual requirements. Laboratories that passed these on-site
evaluations were awarded contracts for analytical services. For total C, N, and S, the laboratory that
submitted the most competitive bid was audited, and subsequently awarded a contract.
5.5.4.2.3 Quality assurance/quality control of analytical laboratories ~
The QA/QC procedures used for evaluating the analytical laboratories were described by Papp
et al. (1989). Evaluations of analytical laboratory performance were documented by Byers et al. (1990).
86
-------
Table 5-7. Analytical Variables Measured in the DDRP Soil Survey (USEPA, 1988)
Variable
Description of Variable
MOIST Percent air-dry soil moisture measured at the analytical laboratory and expressed as a percentage on an oven-dry
weight basis. Mineral soils were dried at 105°C, organic soils at 60°C.
SAND Total sand is the portion of the sample with particle diameter between 0.05 mm and 2.0 mm. It was calculated as
the summation of percentages for individual sand fractions: VCOS + COS + MS + FS + VFS.
VCOS Very coarse sand is the sand fraction between 1.0 mm and 2.0 mm. It was determined by sieving the sand which
had been separated from the silt and clay.
COS Coarse sand is the sand fraction between 0.5 mm and 1.0 mm. It was determined by sieving the sand which had
been separated from the silt and clay.
MS . Medium sand is the sand fraction between 0.25 mm and 0.50 mm. It was determined by sieving the sand which
had been separated from the silt and clay.
FS Fine sand is the sand fraction between 0.10 mm and 0.25 mm. It was determined by sieving the sand which had
been separated from the silt and clay.
VFS Very fine sand is the sand fraction between 0.05 mm and 0.10 mm. It was determined by sieving the sand which
had been separated from the silt and clay.
SILT Total silt is the portion of the sample with particle diameter between 0.002 mm and 0.05 mm. It was calculated by
subtracting from 100 percent the sum of the total sand and clay.
COSI Coarse silt is the silt fraction between 0.02 mm and 0.05 mm. It was calculated by subtracting the fine silt fraction
from the total silt.
FSI Fine silt is the silt fraction between 0.002 mm and 0.02 mm. It was determined by the pipet method (USDA/SCS,
1984) and was calculated by subtracting the clay fraction from the less than 0.02 mm fraction.
CLAY Total clay is the portion of the sample with particle diameter of less than 0.002 mm and is determined using the
pipet method.
PH_H20 pH determined in a deionized water extract using a 1:1 mineral soil to solution ratio and 1:5 organic soil to
solution ratio. The pH was measured with a pH meter and combination electrode.
PH_002M pH determined in a 0.002M calcium chloride extract using a 1:2 mineral soil to solution ratio and 1:10 organic soil
to solution ratio. The pH was measured with a pH meter and combination electrode.
PH_01M pH determined in a 0.01 M calcium chloride extract using a 1:1 mineral soil to solution ratio and 1:5 organic soil to
solution ratio. The pH was measured with a pH meter and combination electrode.
CA_CL Exchangeable calcium determined with an unbuffered 1M ammonium chloride solution. A 1:26 mineral soil to
solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively
coupled plasma atomic emission spectrometry was specified.
MG_CL Exchangeable magnesium determined with an unbuffered 1M ammonium chloride solution. A 1:26 mineral soil to
solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively
coupled plasma atomic emission spectrometry was specified.
K_CL Exchangeable potassium determined with an unbuffered 1M ammonium chloride solution. A 1:26 mineral soil to
solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry was specified.
(Continued)
87
-------
Table 5-7. Continued
Variable
Description of Variable
NA_CL Exchangeable sodium determined with an unbuffered 1M ammonium chloride solution. A 1:26 mineral soil to
solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively
coupled plasma atomic emission spectrometry was specified.
AL_CL Exchangeable aluminum determined with an unbuffered 1M ammonium chloride solution; approximately 1:13
mineral soil solution ratio or 1:52 organic soil to solution ratio are used; inductively coupled plasma atomic
emission spectrometry is specified.
CA_OAC Exchangeable calcium determined with 1M ammonium acetate solution buffered at pH 7.0. A 1:26 mineral soil to
solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively
coupled plasma atomic emission spectrometry was specified.
MG_OAC Exchangeable magnesium determined with 1M ammonium acetate solution buffered at pH 7.0. A 1:26 mineral
soil to solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry or
Inductively coupled plasma atomic emission spectrometry was specified.
K_OAC Exchangeable potassium determined with 1M ammonium acetate solution buffered at pH 7.0. A 1:26 mineral soil
to solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry was specified.
NA_OAC Exchangeable sodium determined with 1M ammonium acetate solution buffered at pH 7.0. A 1:26 mineral soil to
solution ratio and 1:52 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively
coupled plasma atomic emission spectrometry was specified.
CEC_CL Cation exchange capacity determined with an unbuffered 1M ammonium chloride solution is the effective CEC
which occurs at approximately the field pH when combined with the acidity component. A 1:26 mineral soil to
solution ratio and 1:52 organic soil to solution ratio were used. Samples were analyzed for ammonium content by
one of three methods: automated distillation/titration; manual distillation/automated titration; or ammonium
displacement/flow injection analysis.
CEC_OAC Cation exchange capacity determined with 1M ammonium acetate solution buffered at pH 7.0 is the theoretical
estimate of the maximum potential CEC for a specific soil when combined with the acidity component. A 1:26
mineral soil to solution ratio and 1:52 organic soil to solution ratio were used. Samples were analyzed for
ammonium content by one of three methods: automated distillation/titration; manual distillation/automated
titration; or ammonium displacement/flow injection analysis.
ACJ3ACL Total exchangeable acidity determined by titration in a buffered (pH 8.2) barium chloride triethanolamine
extraction using a 1:30 soil to solution ratio.
CA_CL2 Extractable calcium determined by a 0.002M calcium chloride extraction. A 1:2 mineral soil to solution ratio and
1:10 organic soil to solution ratio were used. The calcium is used to calculate lime potential. Atomic absorption
spectrometry or inductively coupled plasma atomic emission spectrometry was specified.
MG_CL2 Extractable magnesium determined by a 0.002M calcium chloride extraction. A 1:2 mineral soil to solution ratio
and 1:10 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively coupled plasma
atomic emission spectrometry was specified.
K_CL2 Extractable potassium determined by a 0.002M calcium chloride extraction. A 1:2 mineral soil to solution ratio
and 1:10 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively coupled plasma
atomic emission spectrometry was specified.
NA_CL2 Extractable sodium determined by a 0.002M calcium chloride extraction. A 1:2 mineral soil to solution ratio and
1:10 organic soil to solution ratio were used. Atomic absorption spectrometry or inductively coupled plasma
atomic emission spectrometry was specified.
(Continued)
88
-------
Table 5-7. Continued
Variable
Description of Variable
FE_CL2 Bctractable iron determined by a 0.002M calcium chloride extraction. A 1:2 mineral soil to solution ratio and 1:10
organic soil to solution ratio were used. Atomic absorption spectrometry or inductively coupled plasma atomic
emission spectrometry was specified.
AL_CL2 Extractable aluminum determined by a0.002M calcium chloride extraction. A 1:2 mineral soil to solution ratio and
1:10 organic soil to solution ratio were used. The aluminum concentration obtained from this procedure is used
to calculate aluminum potential. Atomic absorption spectrometry or inductively coupled plasma atomic emission
spectrometry was specified.
FE_PYP Extractable iron determined by a 0.1 M sodium pyrophosphate extraction using a 1:100 soil to solution ratio, The
pyrophosphate extract estimates organically-bound iron. Atomic absorption spectrometry or inductively coupled
plasma atomic emission spectrometry was specified.
AL_PYP Extractable aluminum determined by a 0.1 M sodium pyrophosphate extraction using a 1:100 soil to solution ratio.
The pyrophosphate extract estimates organically-bound aluminum. Atomic absorption spectrometry or inductively
coupled plasma atomic emission spectrometry was specified.
FE_AO Extractable iron determined by an ammonium oxalate-oxalic acid extraction using a 1:100 soil to solution ratio.
The acid oxalate extract estimates organic and amorphous iron oxides. Atomic absorption spectrometry or
inductively coupled plasma atomic emission spectrometry was specified.
AL_AO Extractable aluminum determined by an ammonium oxalate-oxalic acid extraction using a 1:100 soil to solution
ratio. The acid oxalate extract estimates organic and amorphous aluminum oxides. Atomic absorption
spectrometry or inductively coupled plasma atomic emission spectrometry was specified.
SI_AO Extractable silicon determined by an ammonium oxalate-oxalic acid extraction using a 1:100 soil to solution ration;
inductively coupled plasma atomic emission spectrometry is specified.
FE_CD Extractable iron determined by a sodium citrate-sodium dithionite extraction using a 1:30 soil to solution ratio.
The citrate dithionite extract estimates non-silicate iron. Atomic absorption spectrometry or inductively coupled
plasma atomic emission spectrometry was specified.
AL_CD Extractable aluminum determined by a sodium citrate-sodium dithionite extraction using a 1:30 soil to solution
ratio. The citrate dithionite extract estimates non-silicate aluminum. Atomic absorption spectrometry or
inductively coupled plasma atomic emission spectrometry was specified.
SO4JH2O Extractable sulfate determined with a double deionized water extract. This extraction approximates the sulfate
which will readily enter the soil solution and uses a 1:20 soil to solution ratio. Ion chromatography was specified.
SO4_PO4 Extractable sulfate determined with a 0.016M sodium phosphate (500 mg P/L) extract. This extraction
approximates the total amount of adsorbed sulfate and uses a 1:20 soil to solution ratio. Ion chromatography
was specified.
SO4_0 Sulfate remaining in a 0 mg S/L. solution following equilibration with a 1:5 mineral soil to solution ratio and 1:20
organic soil to solution ratio. The data are used to develop sulfate isotherms. Ion chromatography was specified.
SO4_2 Sulfate remaining in a 2 mg S/L solution following equilibration with a 1:5 mineral soil to solution ratio and 1:20
organic soil to solution ratio. The data are used to develop sulfate isotherms. Ion chromatography was specified.
(Continued)
89
-------
Table 5-7. Continued
Variable
Description of Variable
SO4_4 Sulfa.te remaining in a 4 mg S/L solution following equilibration with a 1:5 mineral soil to solution ratio and 1:20
organic soil to solution ratio. The data are used to develop sulfate isotherms. Ion chromatography was specified.
SO4JB Sulfate remaining in a 8 mg S/L solution following equilibration with a 1:5 mineral soil to solution ratio and 1:20
organic soil to solution ratio. The data are used to develop sulfate isotherms. Ion chromatography was specified.
SO4J 6 Suifata remaining in a 16 mg S/L solution following equilibration with a 1:5 mineral soil to solution ratio and 1:20
organic soil to solution ratio. The data are used to develop sulfate isotherms. Ion chromatography was specified.
SO4_32 Sulfate remaning in a 32 mg S/L solution following equilibration with a 1:5 mineral soil to solution ratio and 1:20
organic soil to solution ratio. The data are used to develop sulfate isotherms. Ion chromatography was specified.
CJTOT Total carbon determined by rapid oxidation followed by thermal conductivity detection using an automated CHN
analyzer. Total carbon can be used to characterize the amount of organic material in the soil.
N_TOT Total nitrogen determined by rapid oxidation followed by thermal conductivity detection using an automated CHN
analyzer. Total nitrogen can be used to characterize the organic material in the soil.
S_TOT Total sulfur determined by automated sample combustion followed by infrared detection or titration of evolved
sulfur dioxide.
90
-------
A priori measurement quality objectives (MQOs) were established for all analyses performed by
the analytical laboratories (Table 5-8). MQOs are statements of the levels of uncertainly that a data
user is willing to accept for the planned purposes of the data. The wide variety of data uses planned
by the DDRP made it difficult to set user-specific MQOs. The MQOs for the sampling, preparation,
and analysis phases were defined and specified in the QA plan (Papp et al., 1989). The initial
analytical MQOs were established in accordance with the DDRP data users based upon selection of
appropriate methods to obtain the data. The MQOs were reviewed by other persons familiar with
analytical methods and soil characterization techniques, including soil chemists and laboratory
personnel. Modifications to the MQOs and to the protocols were implemented using information
gained from the DDRP NE and SBRP surveys, from peer review comments, or according to the limita-
tions of a particular procedure or analytical instrument. The MQOs were translated into detection limits
and precisions that the analytical laboratories were required to meet (Tables 5-9 and 5-10).
5.5.4.2.3.1 Audits. Each analytical laboratory was audited at least three times. The first audit was
conducted after the PE sample data were evaluated, before the laboratories became operational. The
remaining audits occurred after sample analysis had begun, and included a review of the data from
audit and QC samples (Byers et al., 1990).
5.5.4.2.3.2 Quality control samples. QC samples were created and used by the analytical labora-
tories to maintain random and systematic errors within specified limits (see Figure 5-6, and Tables 5-8,
5-9, and 5-10). They were used to evaluate the calibration and standardization of instruments and to
identify problems such as contamination or analytical interference. QC samples included calibration
blanks, reagent blanks, QC check samples, detection limit QC check samples, matrix spikes, analytical
duplicates, and ion chromatography standards. Failure to meet the specified quality limits could result
in rejection of a batch. Detailed descriptions of the use of QC samples was given by Byers et al.
(1990).
5.5.4.2.3.3 Audit samples. Audit samples (Figure 5-6) differed from QC samples in that they were
submitted as blind samples to the analytical laboratories. These were samples of soils that had been
well characterized before the DDRP analyses began. Field audit samples were processed by the
sampling crews as if they had just been obtained from the excavated soil pit. The preparation labora-
tory inserted three field audit samples and three laboratory audit samples into batches so that their
identities and composition were unknown to the analysts. Thus, data from these samples provided an
independent assessment of data quality and a means for monitoring the QC procedures. As with the
preparation laboratory duplicates and the field duplicates, the audit samples provided a measure of
precision (i.e., standard deviation) that could be compared to the precision MQO's. Table 5-11 sum-
marizes the attainment of precision MQO's in the M-APP, as indicated by data from the laboratory
audit samples. The only variable for which within-batch standard deviation (12 percent) exceeded the
MQO (10 percent) was silicon. This variable was not measured in the previous DDRP surveys, and the
laboratories apparently had some difficulty in the analysis of SI_AO at higher concentrations. A
detailed evaluation of the attainment of MQO's is given by Byers et al (1990).
A unique QA/QC feature for the M-APP was the use of the laboratory audit samples to estimate
analytical bias, and the corresponding field audit samples to estimate system-wide bias. We found
91
-------
Table 5-8. Withln-Batch Precision Objectives for the Analytical Measurement Quality Objectives
(from Byers et al., 1990)
Parameter
MOIST
SANDb
SILTb
CLAYb
PH H20
PH 002M
PH_01M
CA CL
MG CL
K CL
NA CL
AL_CL
CA OAC
MG OAC
K OAC
NAJDAC
CEC CL
CEC OAC
AC_BACL
CA CL2
MG CL2
K CL2
NA CL2
FE CL2
ALJDL2
FE PYP
AL PYP
FE AO
AL~AO
Reporting
units
wt%
II
it
u
pH units
u
u
rneq/100g
u
11
u
H
rneq/100g
H
II
tl
rneq/100g
it
u
rneq/100g
H
tl
H
II
U
Vi/t%
II
M
H
SD
0.3
3.0
3.0
2.0
0.10
0.10
0.10
0.02
0.02
0.02
0.02
0.20
0.02
0.02
0.02
0.02
0.25
0.25
1.0
0.05
0.005
0.005
0.005
0.01
0.05
0.03
0.03
0.03
0.03
Mineral8
%RSD
2.5
...
—
10
10
10
10
15
10
10
10
10
10
10
15
5
10
10
10
15
15
10
10
10
10
Organic9
Knot
12.0
—
—
0.20
0.20
0.20
0.20
1.3
0.20
0.20
0.20
0.20
2.5
2.5
6.7
1.0
0.05
0.05
0.05
0.07
0.33
0.30
0.30
0.30
0.30
SD
0.3
0.10
0.10
0.10
0.10
0.10
0.10
0.10
0.50
0.10
0.10
0.10
0.10
0.25
0.25
1.0
0.20
0.05
0.025
0.025
0.05
0.20
0.03
0.03
0.03
0.03
%RSD
2.5
—
...
10
10
10
10
15
10
10
10
10
10
10
15
5
10
10
10
15
15
10
10
10
10
Knot
12.0
...
. ...
1.0
1.0
1.0
1.0
3.3
1.0
1.0
1.0
1.0
2.5
2.5
6.7
4.0
0.50
0.25
0.25
0.33
1.3
0.30
0.30
0.30
0.30
(Continued)
0 Withln-faatch precision by standard deviation below the knot and by percent relative standard deviation above
the knot for mineral and organic samples, respectively.
b Particle size and sulfate isotherm parameters were not analyzed for organic samples.
NOTE: For the preparation samples, lower tier objective is same as value in table, upper tier objective is 1.5x
value: for the field samples, lower tier objective is 2x value, upper tier objective is 3x value (Papp et al., 1989).
92
-------
Table 5-8. (Continued)
Reporting
Parameter units
SI AO
FE CD
AL CD
SO4 H2O mg S/kg
SO4_PO4
SO4_0-32b mg S/L
C TOT wt %
N TOT
S TOT
Mineral8
SD
0.03
0.03
0.03
1.5
1.5
0.10
0.05
0.015
0.002
%RSD
10
10
10
10
10
5
10
10
10
Knot
0.30
0.30
0.30
15.0
15.0
2.0
0.50
0.15
0.02
SD
0.03
0.03
0.03
1.5
1.5
0.05
0.015
0.002
Orqanica
%RSD
10
10
10
10
10
10
10
10
Knot
0.30
0.30
0.30
15.0
15.0
0.50
0.15
0.02
a Within-batch precision by standard deviation below the knot and by percent relative standard deviation above
the knot for mineral and organic samples, respectively.
b Particle size and sulfate isotherm parameters were not analyzed for organic samples.
NOTE: For the preparation samples, lower tier objective is same as value in table, upper tier objective is 1.5x
value; for the field samples, lower tier objective is 2x value, upper tier objective is 3x value (Papp et al., 1989).
93
-------
Table 5-9. Contract-Required Detection Limits for the Analytical
Laboratories (From Byers et al., 1990)
Reporting CRDLa
Parameter
CA CL
MG CL
K CL
NA CL
AL_CL
CA OAC
MG OAC
K OAC
NA_OAC
CEC_CL
H
CEC_OAC
H
AC_BACL
11
CA CL2
MG Cl_2
K CL2
NA CL2
FE CL2
AL_CL2
FE PYP
AL PYP
FE AO
AL AO
SI AO
FE CD
AL CD
units Units
meq/100g 0.0033
0.0054
0.0017
0.0029
0.0145
meq/100g 0.0033
0.0054
0.0017
0.0029
meq/100g 0.1500b
0.0750°
0.1500b
0.0750°
0.2500d
1 .2500d
meq/100g 0.0050
0.0008
0.0003
0.0004
0.0005
0.001 1
Wt % 0.0050
0.0050
0.0050
0.0050
0.0050
0.0020
0.0020
mg/L
0.05
0.05
0.05
0.05
0.10
0.05
0.05
0.05
0.05
0.0075b
1.05°
0.0075b
1.05°
0.005d
0.025d
0.5
0.05
0.05
0.05
0.10
0.10
0.50e
0.50e
0.50e
0.50e
0.50s
0.50e
0.50e
(Continued)
0 Contract-required detection limit in reporting units and instrument units (mg/L or ppm),
respectively.
For titration method, in meq/100g and meq, respectively.
0 For flow injection analysis (FIA) method, in meq/100g and meq/L,
respectively.
d For mineral samples and organic samples, in meq/100g and meq, respectively.
8 DetecCon limit in parts per million.
NOTE: Detection limits not applicable for the physical parameters, soil pH, and the SO4_2-32
parameters.
94
-------
Table 5-9. (Continued)
Parameter
SO4 H2O
SO4 PO4
SO4_0
C TOT
N TOT
S TOT
Reporting
units
mg S/kg
11
mg S/L
wt%
it
II
CRDLa
Units
0.5000
0.5000
0.0250
0.0100
0.0050
0.0010
mg/L
0.025e
0.0259
0.025
100.0e
50.0e
10.0e
a Contract-required detection limit in reporting units and instrument units (mg/L or ppm),
respectively.
b For titration method, in meq/100g and meq, respectively.
0 For flow injection analysis (FIA) method, in meq/100g and meq/L,
respectively.
d For mineral samples and organic samples, in meq/100g and meq, respectively.
8 Detection limit in pahs per million.
NOTE: Detection limits not applicable for the physical parameters, soil pH, and the SO4_2-32
parameters.
95
-------
Table 5-10. Detection Limits for Evaluation of Contractual Compliance and for Independent
Assessment of Analytical and System-Wide Measurement (From Byers et al.,
1990).
Contractual compliance8
Independent assessment13
Parameter
CA OAC
MCfOAC
K_OAC
NAJDAC
CEC CLd
CEC~CL°
CEC~OACd
CECTOAC9
AC BACL1
AC_BACL9
CA_CL2
MG CL2
K CL2
NA CL2
FE~CL2
AL~CL2
FE PYP
AL~PYP
FE AO
AL AO
Sl"AO
FE_CD
AL_CD
SO4 H2O
SO4~PO4
SO4J)
CTOT
N~TOT
S~TOT
Reporting
units
meq/1 OOg
meq/100g
meq/1 OOg
wt %
mg S/kg
K
mg S/L
wt %
CRDL
IDL
0.003
0.005
0.002
0.003
0.15
0.075
0.15
0.075
0.25
1.25
0.0050
0.0008
0.0003
0.0004
0.0005
0.0011
0.005
0.005
0.005
0.005
0.005
0.002
0.002
0.50
0.50
0.025
0.010
0.005
0.001
0.0022
0.0016
0.0009
0.0010
0.0530
0.0175
0.0629
0.0224
0.1738
0.6240
0.0014
0.0004
0.0001
0.0002
0.0004
0.0009
0.0022
0.0029
0.0023
0.0036
0.0026
0.0009
0.0011
0.3339
0.3923
0.0167
0.0056
0.0015
0.0007
IIDL
0.0035
0.0024
0.0016
0.0026
.0.1043
0.0116
0.0785
0.0055
0.5942
2.2500
0.0003
0.0001
0.0004
0.0009
0.0031
0.0030
0.0025
0.0020
0.0026
0.0023
0.0007
0.0008
0.4001
0.7552
0.0350
0.0059
0.0016
0.0021
SDL %RS>SDLC
CA CL
MG~ CL
KCL
NA CL
AL CL
meq/1 OOg
H
H
a
u
0.003
0.005
0.002
0.003
0.015
0.0026
0.0015
0.0009
0.0012
0.0111
0.0032
0.0018
0.0012
0.0027
0.0098
0.1129
0.0251
0.0180
0.0130
0.3161
64.9*
82.6
99.2
40.5*
93.1
0.0548
0.0239
0.0291
0.0105
0.1985
0.2736
0.4513
0.2060
4.3067
0.0861
0.0140
0.0044
0.0034
0.0026
0.0126
0.0416
"0.0502
0.1340
0.0637
0.0395
0.0990
0.0272
2.0640
3.7879
0.2901
0.0656
0.0022
0.0077
76.3*
83.8
96.6
59.4*
100.0
100.0
100.0
100.0
88.3
100.0
87.0
98.5
95.9
24.8*
80.5
95.1
94.0
78.9*
94.1
19.2*
99.2
99.4
95.7
97.6
88.9
96.8
99.5
67.1*
Contract-required detective limit and actual reported instrument detection limit, respectively, in reporting units; based on
instrument analysis of calibration blanks.
Independent instrument detection limit and system detection limit, respectively, in reporting units; based on analysis
of DL-QCCS and FAL samples, respectively (Figure 5-6), calculated independently from the contractual requirements.
Percent of routine samples with values exceeding the system detection limit (asterisk denotes fewer than 80 percent).
Detection limits for the titration method.
Detection limits for the flow injection analysis method.
Reported for mineral samples.
Reported for organic samples.
•NOTE: Detection limits not applicable for physical parameters, pH, and the remaining sulfate isotherm parameters.
96
-------
MINERAL SOZI. BATCHES
SAMPLING
PREPARATIOH
ANALYSIS
Routine
sample
PD
Routine
PD
Routine
I
FD
=•0
1
PD
FD
QE
sample!
FAL
(C)
r-T
Fl
(<
PD
J~
F
(
— IFAP' — 1
(B/Bw)
pair
n_.
u,
:>
,— 'FAP' — ,
(B/Bw)
pair
S_
M,
= >
i— IF!
(B
P
1 1 o<
] no preparation 1
f—l 1 — • i — ILAP' — i
LAI. (B/Bw)
(C) pair
ijpl —
/Bw)
iir
r-*1-! r-1^
LAI. (B
(C) pa
/Bw)
Lr A
: samples — j
OCAS
<*)
—II—1 "-I
OCAS
D
3 FD/Routin« 3
PD/Routine FD/PD
PAP
OCAS
FD - iield duplicate; FAL - low-range Hold nudLt; FAP - field audit pair; PD - preparation duplicate; LAI. -
low-range laboratory audit; LAP - laboratory audit pair; QCAS - quality control audit oample; AD - analytical
duplicate.
ORGANIC SOIL BATCHES
QE oamplea
no preparation
PREPARATION
ANALTSIS
PD
PD
Routine
Routine
FD
FD
PD
PD
(C)
three
i — ^r'
tn'
«>'
>)
roe
(0)
three
1 — ^
(<
th
^O' —
3)
ree
PD/Routine
FD/Routine
FD/PD
|— QC (ample* —)
AD
QCAS
(Oa)
QCAS
FD - tiold dupliCAte; FAO « field audit triplicate (organic); PD - preparation duplicate; LAO - laboratory
audit triplicate (organic); QCAS - quality control audit sample; AD • analytical duplicate.
Eyeten within-batch pro
Within-batch precinion.
Hithin-batch precision.
Syotem detection limit.
Syotem detection limit.
System accuracy and within-/b«tween-batch precision.
Analytical accuracy.
Analytical accuracy and within-/between-batch precision.
Within-batch precision.
Analytical accuracy.
Figure 5-6. Quality assurance and quality control soil samples for mineral and organic batches in
the M-APP Region (from Byers et al., 1990).
97
-------
Table 5-11. Attainment of Measurement Quality Objectives for Precision by
the Analytical Laboratories as Determined from Blind Audit
Samples for the M-APP Region (from Byers et al., 1990)
Varicible
SAND
SILT
CLAY
PH H20
PH 002M
PH_01M
CA CL
MG CL
K CL
NA CL
AL_CL
CA OAC
MG OAC
K OAC
NA_OAC
CEC CL
CEC OAC
AC_BACL
CA CL2
MG CL2
K CL2
NA CL2
FE CL2
AL_CL2
FE PYP
AL_PYP
Notes: Y
N
S
N/A
N/D
Attainment
Lower limit
Y
Y
Y
Y
Y
Y
Y
Y
Y
Y
N/D
Y
Y
Y
Y
N/D
N/D
N/D
Y
S
Y
Y
Y
Y
N/D
N/D
= Met MQO.
= Did not meet MQO.
= Slightly exceeded MQO.
= Not applicable because no MQO set.
= No data.
of MQO
Upper limit
N/A
N/A
N/A
N/A
N/A
N/A
Y
Y
Y
Y
S
Y
Y
Y
Y
Y
Y
Y
N/D
Y
N/D
Y
N/D
N/D
Y
Y
(Continued)
98
-------
Table 5-11. (Continued)
Variable
Attainment of MQO
Lower limit
Upper limit
FE AO
AL AO
SI AO
FE CD
AL_CD
SO4 H20
SO4 PO4
SO4JD
SO4 2
SO4 4
SO4 8
SO4J6
SO4_32
C TOT
N TOT
S-TOT
Notes: Y
N
S
N/A
N/D
N/D
N/D
Y
N/D
N/D
S
N/D
Y
Y
N/D
N/D
N/D
N/D
N/D
Y
Y
= Met MQO.
= Did not meet MQO.
= Slightly exceeded MQO.
= Not applicable because no MQO set.
= No data.
Y
Y
N
Y
Y
Y
Y
Y
Y
Y
Y
Y
Y
Y
Y
Y
99
-------
(Byers et al., 1990) that the field and preparation components of system-wide bias were negligible and
that bias was principally an analytical laboratory component. Table 5-12 shows the bias values
expressed in reporting units. Bias was not determined for the organic audit samples because of the
low number of organic batches analyzed.
5.5.5 Database Management
The mapping data were of two fundamental types: alphanumeric (attribute) and geographic
(map). The alphanumeric data were tabulated on personal computers using dBase III and PC SAS
software systems. All tabular data were incorporated into a series of SAS files on mainframe
computers. The map data were digitized and stored as ARC/INFO files (Section 5.4). The field
sampling and sample preparation data were entered into separate SAS-AF raw data files on personal
computers at EMSL-LV.
An important innovation for the M-APP was the development of a Laboratory Entry and
Verification Information System (LEVIS) that could be utilized at the analytical laboratories. The LEVIS
program facilitated the entry, editing, and review of intermediate data and the calculation of final data
values. The program also performed verification checks for the measurement quality samples and
produced QC summary reports.
After all samples in a batch were analyzed for a given variable, the results were recorded on
raw data forms and subsequently entered into LEVIS by the analytical laboratories. A single entry
system was employed in conjunction with a final visual data check performed by the laboratory
manager prior to formal data submission. The computer system provided each analytical laboratory
with standardized data forms that allowed entry of sample weights, solution volumes, dilution factors,
instrument readings, titrant volumes, and titer acid normalities for routine sample analysis. Separate
data forms were provided for the QC sample data, including reagent blanks, calibration blanks, check
samples, matrix spike analyses, spike solutions, replicate analysis, and calculated detection limits. A
page was also provided to allow documentation of comments by the laboratory analysts or manager.
All raw data forms were archived for possible use as a future reference base (Byers et al., 1990).
The analytical laboratories and the QA staff at EMSL-LV used LEVIS to evaluate data quality. The
analytical laboratories evaluated data as they were produced, and the EMSL-LV QA staff evaluated
preliminary and final data. Preliminary databases contained incomplete data that were amended or
added to by the analytical laboratory and transmitted to the QA staff via modem. The laboratories
delivered the official analytical data to the EMSL-LV QA staff on floppy disks in the LEVIS database
structure. The EMSL-LV QA staff performed additional verification checks according to the criteria
specified by USEPA (1988). Raw data verification was accomplished by a systematic evaluation of
completeness, precision, internal consistency (including the relationships in Table 5-13), and coding
accuracy. Apparent discrepancies were appended with data qualifiers, or flags, unless they could be
corrected. After verification was completed, the databases were "frozen" from further editing by the
EMSL-LV QA staff (Byers et al., 1990). The field and preparation data were verified according to the
criteria specified by Papp et al. (1989, Appendix G and H, respectively).
100
-------
Table 5-12. Analytical Bias Estimates for Mineral Samples (from Byers et al., 1990)
Parameter
SAND
VCOS
COS
MS
FS
VFS
SILT
COSI
FSI
CLAY
PH H20
PH 002M
PH_01M
CA CL
MG CL
K CL
NA CL
AL_CL
CA OAC
MG OAC
K OAC
NA_OAC
CEC CLb
CEC CLC
CEC OACb
CEC OACC
AC_BACL
CA CL2
MG CL2
K CL2
NA CL2
FE CL2
AL CL2
Reporting
units
wt%
it
n
II
it
ll
u
n
11
u
pH units
n
u
meq/100g
n
it
it
u
meq/1 OOg
II
n
tl
meq/1 OOg
meq/1 OOg
U
II
it
meq/1 OOg
11
11
n
u
n
Analytical
bias
0.3211
0.3948
0.7606
0.5632
-0.5363
-0.8188
0.9283
0.2344
0.2929
-0.6168
-0.0319
0.0277
-0.0216
-0.0028
0.0011
0.0010
-0.0013
-0.061 5a
0.0005
0.0007
0.0059
0.0044
-0.0261
0.0217
-0.6132*
0.0523
-0.1611
0.0248
0.0024
0.0005
0.0012
-0.0004
0.0023
a Based on data from B horizon FAP (Figure 5-6) samples only.
b Analysis by titration.
c Analysis by FIA.
d Analyzed by only one laboratory.
(Continued)
Exceeds the system detection limit (SDL).
101
-------
Table 5-12. (Continued)
Parameter
FE PYP
AL_PYP
FE AO
AL AO
SI AO
FE CD
AL_CD
SO4 H2O
SO4 PO4
SO4 0
SO4 2
SO4 4
SO4 8
SO4_16
SO4_32
C TOT*1
N TOTd
s'roT'1
Reporting
units
wt%
II
ll
ll
ll
it
it
mg S/kg
n
mg S/L
u
ll
it
ti
ll
wt%
n
n
Analytical
bias
-0.0183
-0.0043
0.0099
0.0009
-0.0034a
0.0136
0.0030
0.1055
2.0341
0.0358
0.0446
0.0526
0.0546
0.1053
0.1026
-0.0352
0.0016
-0.0007
0 Based on data from B horizon FAP (Figure 5-6) samples only.
b Analysis by titration.
0 Analysis by FIA.
d Analyzed by only one laboratory.
Exceeds the system detection limit (SDL).
102
-------
Table 5-13. Soil Chemistry Relationships and Delimiters Used by LEVIS
Soil chemistry
relationship
Delimiter
PH_H20 > PH_002M > PH_01M
CA_CL + MG_CL + K_CL + NA_CL < CEC_CL
AL_CL < 0.01 meq/100g
CA_OAC + MG_OAC + KJDAC + NA_OAC
< CECJDAC
CEC_OAC > CEC_CL
CEC_OAC 4- CLAY < 50
CEC_CL 4- CLAY < 50
MG_CL2 < MG_CL * 1.10
K_CL2 < K_CL* 1.10
NA_CL2 < NA_CL* 1.10
4 < SO4_H2O 4- SO4_0 < 20
S/kg
SO4_PO4 > SO4_H20a
SO4_32 > SO4_16 > SO4_8 > SO4_4 > SO4_2
> SO4_0
SO4_32N > SO4_16N > SO4_8N > SO4_4N >
SO4_2N > SO4_ON
7 < C_TOT 4- N_TOT < 50
40 < C TOT 4- S TOT < 400
if pH differences were > 0.05 units
if CEC_CL > 1 meq/100g and pH_H2O < 7
if PHJH20 > 6.0
if CEC_CL > 1 meq/100g and pH_H2O < 7
if CEC_CL > 1 meq/100g and PH_H2O < 7
if CLAY > 1.0 and CECJDL > 1 meq/100g
if CLAY > 1.0 and CEC_CL > 1 meq/100g
if MG_CL2 > 0.008 meq/100g
if K_CL2 > 0.003 meq/100g
if NA_CL2 > 0.03 meq/100g
if SO4_H2O > 2 mg S/kg and SO4_0 > 0.1 mg
if SO4_PO4 > 1 mg S/kg
if SO4_32N < 7.5 mg S/kg
b if SO4_32N > 7.5 mg S/kg
if C_TOT-^ 0.1% and N_TOT > 0.03%
if C TOT > 0.1% and S TOT > 0.005%
a SO4_PO4 was multiplied by 1.10 for organic soils
b N = mean spike concentration of initial solution — (instrument reading x dilution factor)
103
-------
Verified data from the analyses conducted during sample preparation were later appended to
the verified analytical database. The "frozen" verified field and analytical databases were then sent to
Oak Ridge National Laboratory (ORNL) to undergo validation in cooperation with ERL-C and EMSL-LV
personnel. The validation procedures included a specific assessment of outlying data points for
inclusion or omfssion in validated datasets based on assigned levels of confidence. These outliers
warrant special attention or caution by the data user during analysis of the data. After all data were
evaluated and the suspect values were confirmed or flagged, the validated databases were frozen.
5.5.6 Data Summary
5.5.6.1 Summary of Sampling Class Data
The percentage area of each sampling class in the target population was calculated using the
procedure shown in Figure 5-7. First, the area of each sampling class on each watershed was esti-
mated from the area and composition of each map unit. The regional area of a sampling class was
estimated as a weighted sum over watersheds, using the inverse of the watershed inclusion probability
as a watershed weight. Total area was calculated by summing over all sampling classes. Percen-
tages were calculated by dividing the area of each sampling class by the total area. This procedure
yielded unbiased estimates (Figure 5-8) of the relative areas of sampling classes in the target
population; that is, all watersheds in the regions that meet the conditions stated in Section 5.2.
Depending on the intended use, data from individual soil samples were aggregated to horizons,
pedons, sampling classes, and watersheds (Sections 8.8.8, 9.2.2.3 of Church et al., 1989). Figure 5-9
shows the pedon-aggregated values of pH (water, 0.01 M CaCI2), CEC (NH4CI), base saturation, clay
content, extractable sulfate (water, PO4), and the slope and x-intercept of the sulfate isotherms of the
B-horizons for all pedons sampled using the M-APP sampling class scheme, excluding one apparently
contaminated pedon. B-horizon data were used for the isotherm parameters because sulfate iso-
therms were not determined for 0 horizons, and because the B-horizons account for the bulk of the
sulfate adsorption.
5.5.6.2 Cumulative Distribution Functions
Cumulative distribution functions (CDFs) of the variables included in Figure 5-9 were calculated
for the target population using the procedure shown in Figure 5°-10. Sampling class means were given
weights equal to the percentage of the area of the target population occupied by the corresponding
class. CDFs for the M-APP (Figure 5-11) were obtained by ordering the sampling class means and
summing the weights. Table 5-14 shows medians of these variables for the M-APP.
5.6 DEPOSITION DATA
The regional nature of the Project required estimates of precipitation and atmospheric deposition
(wet and dry) developed in a standardized manner across the eastern United States. Study sites for
the DDRP were selected statistically (Section 5.2) and had no direct information for deposition.
104
-------
:y:: ;;::,;;:' :;:';•:::: 7
'*" - Designate sampling class - ' ,4
- ?..£-" x -v - " "- "/
NO
NO
NO
I
Designate a watershed in
region or subregion
I
Designate a map unit
on watershed
I
Area of map unit
on watershed
1
Area of sampling class
on watershed: Asw
I
Assign weight Ww, equal to inverse
of watershed inclusion probability
YES
Calculate sampling class a'ea (As)
by weighted sum over watersheds
As = £Ww Asw
YES
Calculate total area by summing
over sampling classes Ar = £As
All sampling
classes
designated?
Percent of sampling
class in map unit
•• : jv 7
Calculate percentage areas /
AS/AT /
Figure 5-7. Procedure for calculating regional area in each sampling class.
105
-------
0.4
P
e 0.3
r
c
e
n
t
a
9 0.2
e
o
f
A
r
e
a
0.1
E3 B B
M M N
D K D
B
X
P
B
X
W
C
F
P
H
S
T
T
P
D
T T T
W W X
D M P
T
X
W
Sampling Class
Figure 5-8. Relative areas of sampling classes in soils of the DDRP target population in the
M-APP Region. See Figure 5-3 for identification of sampling classes.
106
-------
5.2
5.0-
CC
4-8
^ 4.6-1
3:
4.4-
4.2
m m m tn
>< 5 u. § 3 w o
m g O g u. I h
SAMPLING CLASS
o
5
d
£
4.0-1
3.8-
3.6
§
SAMPLING CLASS
18
16
•7" 14-
f 12-
o
S, 10"
O
g 8
T 1 1 1 1 1 1
& 8 S: i 5 te i
m g o g a: i h
SAMPLING CLASS
50
40-
so •
< 20 •
w
UJ
W
SAMPLING CLASS
30
20-
>E
o
SAMPLING CLASS
Figure 5-9. Aggregated soil variables for individual pedons in the M-APP Region. See Figure
5-3 for identification of sampling classes (Page 1 of 2).
107
-------
200
?150
tr
}i!ioo-]
0
z
- soH
SAMPLING CLASS
a. so
^
w
E
1 20
8
"1 1 1 T"
i § |
m m ca
SAMPLING CLASS
300
200-
2 100
T
w so-
BMlBMiBN[BN{BXIBXVCFICM!FLVHS^TPtTWTW^TXFTXW
SAMPLING CLASS
700
*[_, 600-
§.500
400-
300-
SAMPLING^ CLASS
a- S
^ i
Figure 5-9. Aggregated soil variables for individual pedons in the M-APP Region. See Figure
5-3 for identification of sampling classes. (Page 2 of 2).
108
-------
Enter
A -
^Designate a sampling class'" ,
Designate pedon
Calculate aggregated soil
variable for pedon
All pedons
in sampling class
Designated?
YES
NO
Calculate mean of soil
variable for sampling class
Assign weight as percentage area
of class in target population
NO
All sampling
classes designated?
Order sampling class means
from lowest to highest
i
Calculate CDF by accumulating ?
sampling class weights
, ,,,,- ,., ,„„,.. ,„
^7
Exit
Figure 5-10. Procedure for calculation of cumulative distribution function for a soil variable in
the M-APP Region.
109
-------
4.8
pH IN WATER
O
3.8 4.0 4.2 4.4
pH IN 0.01 m CALCIUM CHLORIDE
4.6
1.0
f= 0.8
cc
g _,
O 0.6-
a.
tu
> 0.4
O
0.0
6 8 10 12
CEC (Ammonium Chloride (meq/100g)
14
1.0
F 0.8-
DC
o
a.
O 0.64
a.
HI
> 0.4-
0.2-
0.0
0
10 15 20 25 30
% BASE SATURATION
35
40
1.0
F 0.8
DC
O
a.
O 0.6
DC
a.
UJ
> 0.4
0.2
o
0.0
10
15
% CLAY
20
25
30
Figure 5-11. Cumulative distribution functions for pedon-aggregated soil variables for the
Wl-APP Region. (Page 1 of 2).
110
-------
1.0
p 0.8-
CC
O
O 0.6-
cc
Q-
UJ
> 0.4-
0.2-
O
0.0
20 40 60 80
WATER EXTRACTABLE SULFATE (mg/kg)
100
1.0
O
P 0.8
CC
O
Q.
O 0.6
CC
Q.
LU
> 0.4
0.2-
O
0.0
0 50 100 150 200
PHOSPHATE EXTRACTABLE SULFATE (mg/kg)
1.0
o.o
0 100 200 300
X - INTERCEPT OF SULFATE ISOTHERM
P 0.8-
CC.
O
Q.
O 0.6-
o:
Q.
UJ
> 0.4-
0.2-
0.0
10 20 30
SLOPE OF SULFATE ISOTHERM
40
Figure 5-11. Cumulative distribution functions for pedon-aggregated soil variables for the
M-APP Region. (Page 2 of 2).
111
-------
Table 5-14. Medians of Pedon-Aggregated Values of Soil Variables
for the M-App Region
Variable
Units
Median
pH (water)
pH (CaCI2)
CEC
BS
Clay
SO4 (water)
S04 (P04)
Isotherm slope
Isotherm intercept
—
—
meq/100g
%
%
mg/kg
mg/kg
—
mg/kg
4.6
3.9
8.1
9.1
16.7
14.2
52.4
3.7
207.
112
-------
Furthermore, time and budgetary constraints precluded the instrumentation of sites and, thus,
the direct acquisition of any deposition data. As discussed in Sections 2, 3, and 4, the DDRP was
designed to focus on the long-term effects on surface water chemistry of deposition of sulfur.
Although sulfur is the primary deposition variable of interest, complete deposition chemistry is required
for the Level II base cation analyses (Section 9.3), and the Level III watershed modelling (Section 10).
5.6.1 Deposition Scenarios
The major question driving the DDRP concerns the response of surface water chemistry to
atmospheric deposition in the future. For the DDRP studies in the M-APP, we were requested by the
U.S. EPA Office of Air to evaluate three sulfur deposition scenarios. The first deposition scenario was
constant deposition at current levels. The second scenario was current deposition for 5 years,
foHowed by a ramp decrease in sulfur deposition for 10 years to a level 50 percent below current,
which was then maintained for the duration of the modelling simulations. The third scenario was the
same as the second to year 15, then sulfur deposition was ramped upward to a level 20 percent below
current values at 50 years and maintained at that level for simulations longer than 50 years. These
scenarios are illustrated in Figure 5-12.
5.6.2 Temporal Resolution
Current deposition was of interest to the DDRP NE and SBRP regional analyses because of the
Level I analyses conducted to determine (1) the current status of sulfur retention within watersheds
(Section 7) and (2) the current relationship among atmospheric deposition, watershed and soil factors,
and surface water chemistry (Section 8). This interest/requirement led to the development of a
deposition dataset that represented atmospheric deposition as of the early to mid-1980s. This
deposition dataset, the "Long-Term Annual Average" (LTA) dataset, was described fully by Church et
al. (1989).
The watershed models used in the Level III analyses required a fine time resolution of precipi-
tation data for the calibration of their hydrologic modules. This requirement necessitated the develop-
ment of a deposition dataset with a daily resolution of precipitation and a monthly resolution of depo-
sition. In our work for the NE and SBRP, we also used this dataset as a comparative check against
the LTA dataset in (1) the Level I analyses for sulfur retention (Section 7, Church et al., 1989) and
(2) the Level II analyses for sulfate adsorption (Section 9.2, Church et al., 1989) and base cation
depletion (Section 9.3, Church et al., 1989). That comparative work indicated negligible differences in
results obtained using the LTA and Typical Year (TY) datasets. Therefore, we developed and applied
only the TY dataset for the M-APP studies. This was required for the Level III analyses and was
sufficient for the Level II M-APP analyses.
5.6.3 Data Acquisition/Generation
Where possible, we attempted to use deposition data (wet and dry) that were available from
specialized deposition projects within the National Acid Precipitation Assessment Program (NAPAP). A
very difficult constraint of the DDRP analyses, however, was that the datasets used had to be complete
113
-------
FUTURE SULFUR DEPOSITION SCENARIOS
DDRP MID-APPALACHIAN REGION
1ZU
X
• H r\n ••
— I 1UU
LL.
DC
ID 80
LJL
a! 60~
£ 40-
'
!j 20"
ULJ
DC
\
\
\
\ . ______________
i x ______________
\
\ s
\
\s . .
Scenario A
Scenario B
Scenario C
0
25 50 75 100 125
YEAR OF SIMULATION
150
Figure 5-12. Sulfur deposition scenarios for the M-APP Region for Level II and III analyses.
114
-------
in terms of chemical composition (i.e., all major ions), regional coverage, and internal consistency
(e.g., charge balance). Such datasets were not available within NAPAP. Thus, as explained below, we
had to generate such data ourselves as best possible. In the course of this data generation, we
consulted at length with available authorities (both within and external to NAPAP) regarding the
reasonableness of our assumptions, our methods, and the data generated.
The TY dataset was designed to (1) provide a daily resolution of precipitation and a monthly
resolution of deposition in order to be consistent with the hydrologic and model time step require-
ments of the Level III models, and (2) to represent a yearly precipitation regime that was, indeed,
"typical" of current climatological conditions for the study regions. The dataset was used repetitively
(i.e., for each year) for the Level III simulations with appropriate adjustments during the increase or
decrease scenarios (Figure 5-12).
5.6.3.1 Wet Deposition
An approach for determining wet deposition data was developed through close consultation with
A. Olsen and his staff who manage the Acid Deposition System database (ADS) at Battelle-Pacific
Northwest Laboratories (PNL). The ADS database is composed of data from all the major wet deposi-
tion monitoring networks in the United States. After the approach was developed, the actual compo-
nent datasets were developed by A. Olsen and his staff.
Initially in the NE and SBRP DDRP studies, we investigated the use of wet deposition data
derived by spatial interpolation (kriging) of deposition monitored at ADS sites. Several factors
immediately acted to dissuade us from this approach. First was the relatively poor spatial coverage by
the ADS sites, which are widely scattered geographically. As a test of interpolation, we kriged wet
sulfate deposition, in and about the area of the Adirondack State Park (New York), and visually
compared the spatial patterns of wet deposition to these sites with the patterns of sulfate flux from
watersheds in the Adirondacks (see Section 7 for a thorough discussion of computation of sulfur
input/output budgets). Previous work had indicated that sulfur inputs probably are in balance with
sulfur outputs in the Adirondacks (Rochelle et al., 1987; Rochelle and Church, 1987). Visual
comparison indicated that the wet input patterns poorly coincided with the output patterns.
As a comparison, we computed wet sulfate inputs by multiplying together wet sulfate concentra-
tion kriged from ADS sites with precipitation kriged from the much denser network of sites of the
National Oceanic and Atmospheric Administration (NOAA) National Climatic Data Center (NCDC).
Patterns produced by this procedure were in much closer agreement to observed patterns of sulfate
outflux from the Adirondack watersheds.
A second important consideration was the efficacy of interpolating monthly values of deposition
or wet concentration from ADS sites. The geographic sparseness of the ADS network and the occa-
sional paucity of monthly data (e.g., during extremely dry months or months during which samples
were not acceptable due to contamination) argued strongly against this approach (A. Olsen, pers.
comm.). A third consideration was that daily precipitation data,,,needed as inputs to the hydrologic
models of the Level III analyses, were not available from the ADS sites.
115
-------
As a result, we decided to develop for each individual DDRP study site (1) an appropriate typical
year of wet concentration chemistry obtained from a nearby linked ADS site and (2) a daily precipita-
tion dataset for a nearby linked NCDC site for the same year as selected for the typical year deposition
chemistry for the linked ADS site. Wet deposition at the DDRP site is then the product of the wet
chemistry and precipitation datasets. This type of multiplicative approach (in general) has been
discussed and endorsed by Vong et al. (1989).
Sites for wet deposition chemistry (ADS) and daily precipitation (NCDC) were carefully selected
for each DDRP study watershed based on geographic location, elevation, and terrain. DDRP staff
selected the sites in close coordination with A. Olsen and project cooperators involved in the Level III
modelling who were familiar with the requirements of the models and the need for appropriate link-
ages between the precipitation inputs and hydrologic outputs from the study watersheds. The ADS
and NCDC sites selected for pairing with the DDRP study sites [n the M-APP are shown in Plate 5-2.
5.6.3.1.1 Wet deposition chemistry --
Precipitation chemistry data were obtained from the ADS database. For each ADS site, the
entire history (usually less than five years) of daily or weekly data was obtained. The annual cumula-
tive distribution functions (CDFs) for each individual year were compared with the summary CDFs of
data for all years. The typical year was selected as the year that compared best to all years for sulfate
concentration, nitrate concentration, and precipitation. After the typical year was selected, monthly wet
deposition chemistry was computed using the procedures recommended by the Unified Database
Committee (Olsen et al., 1990). Their quarterly criteria were applied to each month. When monthly
data for the typical year selected did not meet the criteria, an alternate typical year was used.
5.6.3.1.2 Daily precipitation -
The same year chosen as the typical year for deposition chemistry was used as the typical year
for precipitation at the linked NCDC site. In a few cases, precipitation data were not available for the
ADS typical year. In this event, we chose the closest years for which precipitation data were available
with respect to sulfate concentration, nitrate concentration, and precipitation. An additional advantage
of using the NCDC sites was that the long-term data available for these sites allowed us to adjust indi-
vidual years and days of data to a long-term norm for the location. In this case, daily precipitation at
each NCDC site was adjusted using a nearby site with 30-year normal monthly and annual data. Sites
and data were obtained from the NCDC tape TD9641: bMonthly Normals of Temperature, Precipitation,
and Heating and Cooling Degree Days 1951-80.
Daily precipitation during a month was adjusted to match the 30-year normal for the month.
Each daily value was multiplied by the ratio of the 30-year normal for the month and the monthly total
for the typical year selected. This procedure also ensured that the typical year annual total matched
the annual 30-year normal. Information on data completeness and quality for the ADS sites is avail-
able from A. Olsen, ERL-C, U.S. EPA.
116
-------
Plate 5-2. ADS and NCDC sites linked with DDRP study sites for the M-APP Region. The
"concentration zone" indicates to which DDRP sites the appropriate typical year of wet
concentration chemistry from the linked ADS site was applied. The "precipitation zone" indicates
to which DDRP sites the appropriate precipitation (i.e., same year as selected for the linked ADS
site and intersecting concentration zone) from the linked NCDC site was applied. See text for
further description.
117
-------
-------
MID-APPALACHIAN REGION
DDRP Study Site Locations
I-APP Study Area
•Precipitatjon Zone
(see caption)
•Concentration Zone
(see caption)
Precipitation Site
Wet Deposition Site
047a Jasper
064a Leading Ridge
065a Penn Slate
072a Virginia
075a Parsons
I 5 I a Scranton
250a Shenandoah
379a Babcock St Park
-"S
I-/
'^'^^P^W^-^^- "7V~
-i- / ^- *. I x •fKSefelj I & 362343 ID029043 } !
" / s' JCQ29Q02 LI ^P x •l^r / /I
*^£^f)^
r'-'" --? A- ~ ! ^!>l|c|4Q75;at! - 'T + Vl +^<>3"« } ^ V
'"'"*'' ' ''"\T "" • ' -i " 'v- ;V,/ ^064^* \iB036MV.\368570 ^"_^-^>
Vf"i:'.'" -' -' *- tjl "'-„' ,,?,; - "y/^~\ ( ^W2\ ~/^ —^-^ff^^.
'-->
^
\e6272i-
h X
^
\'
::!'. ii"-. ;.v^^.:'iiiJ#i^4^8ap' ,
: j-V:- 7 : ZM4104Hi?g68>7^ , / ^'
,,.L-->'.f . O^TMip^ X..
'"•""-, rVt,.!'., -, w" j-i . 4J2(,^^^d3/164/>xf'",7J;i,n,R /I 1
;--..:-. ,^":.^,^J^«f3W&"^'"j«|M '
^"^^'•da!^^^*fo-.
^.^^•^.ri^i*:-
,r . >, ,i. ";, i"'\ S?8rf^. X46S87I ' / • -'.AX=-T. "'" -•"
vi; ' ^;',„„,;, 1,-'v-^:AcliaM-4'' C"
•'"|1(i ^C;;:" ' , ^\s X
. i lit ' ' ' Uj '-. ' w.'j",'" 1 s.S
ft\
-'^"
»t/A.
\ ~
£fa, .,l,iiJ>Si
-------
-------
5.6.3.2 Dry Deposition
The determination of representative typical year estimates of dry deposition at the DDRP study
sites was a very difficult task. The accurate measurement of dry deposition to watersheds is a devel-
oping art, and for the purposes of the DDRP, no network of sites existed that was able to provide the
regionally consistent information that our analyses required. Instead, we had to rely on estimates from
a variety of modelling and inferential techniques. (We use the term "estimates", as opposed to "data,"
to describe derived values for dry deposition for all variables. To describe the grouping of these
estimates, however, we use the term "database".) Information on estimates of dry sulfur deposition
from the Regional Acid Deposition Model (RADM) (Chang et al., 1987) was obtained from R. Dennis
(AREAL-RTP) and S. Seilkop (Analytical Sciences, Inc.), as was information on possible annual-scale
relationships among fine particle dry deposition of base cations and chloride and wet deposition of
those same ions. We combined this information with other information on (1) dry deposition to
surfaces, (2) canopy scavenging, (3) throughfall, and (4) pertinent information on interactions among
atmospheric deposition and watershed ion budgets to construct complete (major ions) suites of dry
deposition to represent a typical year for each of the DDRP study sites.
5.6.3.2.1 Sulfur--
As for the NE and SBRP study sites (Church et al., 1989), interim or first-stage dry sulfur
deposition estimates for M-APP study sites were based upon output available from RADM (R. Dennis
and S. Seilkop, pers. comm., and Clark et al., 1989). First-stage estimates were based on the
simulation of six three-day episodes and the results were averaged to establish regional dry
deposition. "Ground-truth" data on dry sulfur deposition from a sparse number of measurement sites
(Hicks et al., 1986; Hosker and Womack, 1986) were used to geographically adjust (spatially calibrate)
the RADM output. The RADM model provided output for points on an 80 x 80 km grid. Estimates at
those points were then kriged to individual DDRP study sites. As was done for DDRP sites in the
adjoining NE region [to adjust for potential biases in the RADM estimates; see Church et al. (1989) for
a more detailed discussion], these first-stage estimates were adjusted upwards by 20 percent
(increasing estimated total sulfur deposition by about 9 percent on a regional average). This provided
the second-stage sulfur deposition estimates.
The next step was to apportion the dry sulfur deposition on a monthly basis. Because scav-
enging of dry sulfur deposition should be a function of canopy development, we used the watershed
vegetation information from the DDRP mapping (Section 5.4.1) to adjust for monthly partitioning. This
was a two-step process. First, we assigned a leaf area index (LAI) (Table 5-15) to each vegetation
type (coniferous, deciduous, and open), based, in part, on values used by Goldstein and Gherini
(1984). We used two variations on this approach: (1) we assigned an LAI of 0.25 to deciduous
vegetation during the months of November through March and (2) we partitioned our "mixed" vegeta-
tion class as half deciduous and half coniferous. Second, we applied an iterative predictor-corrector
technique to apportion the monthly deposition so that its sum closely approximated the second-stage
annual dry sulfur deposition totals. Application of these procedures provided a third-stage (final) dry
sulfur deposition dataset for which the annual sum of the monthly dry deposition was within a few
percent of the second-stage annual value on the average for any watershed.
118
-------
Table 5-15. Monthly Values of Leaf Area Index (LAI) Used to Apportion Annual
Dry Deposition to Monthly Values
Month
LAId
LAIc
LAlo
Jan
Feb
Mar
Apr
May
Jun
Jul
Aug
Sep
Oct
Nov
Dec
0.25
0.25
0.25
0.5
1.0
2.5
4.0
4.5
4.5
1.0
0.25
0.25
12
12
12
12
13
14
15
15
15
15
14
13
1
1
1
1
1.5
1.5
1.5
1.5
1.5
1
1
1
LAI d = leaf area index for deciduous vegetation
LAI c — leaf area index for coniferous vegetation
LAI o = leaf area index for open areas
119
-------
5.6.3.2.2 Base cations and chloride --
Computation of individual watershed values for dry deposition of base cations (Ca2+, Mg2+,
Na+, K+) and chloride (Cl~) involved quite a number of considerations and computational steps. At
the heart of the computation was the development of a technique (Eder and Dennis, 1990) that used
regression analysis between measured annual wet deposition and the annual geometric means of
ambient air concentration (used with deposition velocities to compute dry deposition). Data used in
the development of the technique and relationships were obtained from the Acid Precipitation in
Ontario Study (APIOS) of the Ontario Ministry of the Environment. (For a description of the network
and its data collection and analysis techniques, see Chan et al., 1982a,b, and Tang et al., 1986.)
Because of the manner in which the ambient concentrations were measured, these relationships
probably apply only to fine-particle (< 2 ,wm) dry deposition. For computing annual fine-particle dry
deposition, it was assumed that deposition velocities were roughly equivalent among the base cations
at 0.8 cm/sec for heavily forested vegetative situations (Eder and Dennis, 1990). (This is the condition
for the DDRP watersheds, inasmuch as most have vegetative coverage of at least 80 percent.) The
annual fine-particle dry deposition had to be partitioned into monthly components. First, the annual
values were partitioned based upon 13 28-day months (Eder and Dennis, 1990), and then they were
repartitioned by DDRP staff into the 12 months comprising the TY dataset.
Coarse-particle (> 2/
-------
Scavenging of dry deposition by vegetative canopies (especially coniferous canopies) is subject
to a pronounced "edge effect", whereby lower windspeeds and ambient concentrations within interior
canopies result in markedly lower effective dry deposition to those interior regions (Dasch, 1987;
Grennfelt, 1987). We reasoned that this process could be represented by a function of the form of the
well-known Michaelis-Menton equation and used to adjust the "effective" coniferous canopy scaveng-
ing. In this way, scavenging of base cations and chloride within our watersheds would not be com-
puted using total coniferous LAIs, but rather the coniferous LAIs would be adjusted (in effect) so that
as the areal coniferous coverage of a watershed increased, its effect on scavenging reached an effec-
tive plateau rather than increasing linearly. We used this approach to adjust total dry deposition until
the chloride budgets approximately balanced in undisturbed NE sites. The final equation used in the
adjustment was
% CONe = (30 * % CON)/(15 + % CON)
where % CON = mapped percent coniferous coverage
% CONe = effective percent coniferous coverage
(5-1)
These computations of dry base cation deposition leave a great deal to be desired. The final values,
however, relate well to (1) estimates of dry deposition-to-wet deposition ratios observed by Lindberg et
al. (1986) fora southeastern forested catchment, and (2) previously modelled estimates at Woods and
Panther Lakes in the Adirondack Mountains (R. Munson, pers. comm.).
5.6.3.2.3 Nitrate and ammonium --
We had no objective or mechanistic approach for estimating dry deposition of nitrate and
ammonium. Instead, we assumed that total dry deposition of nitrate was equal to wet deposition and
that total dry deposition of ammonium was equal to one-half wet deposition. These ratios approximate
values measured by Lindberg et al. (1986) in an eastern forested watershed.
5.6.3.2.4 Hydrogen ion --
We computed dry H+ deposition as the difference between dry anions and other dry cations.
When the sum of other dry cations was greater than the sum of dry anions, we set dry H+ to zero.
5.6.3.2.5 Ion ratios --
The ratios of dry deposition to wet deposition for all ions for the M-APP study sites for the TY
dataset are shown in Table 5-16 in relation to values obtained for the NE and SBRP.
5.6.3.2.6 Comparisons with direct measurements --
Although extensive data do not exist with which to compare the DDRP estimates, Church et al.
(1989) used some limited information (B. Hicks, pers. comm.) for this purpose for the previous esti-
mates made for the NE and SBRP. We had no comparable information for the M-APP Region.
121
-------
Table 5-16. Ratios of Dry Deposition to Wet Deposition for DDRP Study Sites
for the Typical Year (TY) Deposition Datasets
NE
S042"
Ca2+
Mg2+
Na+
K+
cr
aN03-
aNH4+
H+
M-APP
S042'
Ca2+
Mg2+
Na+
K+
cr
aN03-
aNH4+
H+
Median
0.44
1.13
1.92
1.29
1.56
0.38
1.0
0.5
0.47
Median
0.85
1.21
1.49
1.01
1.17
0.28
1.0
0.5
0.92
Mean
0.48
1.12
1.82
1.29
1.66
0.33
1.0
0.5
0.46
Mean
0.87
1.30
1.43
0.94
1.12
0.27
1.0
0.5
0.94
Standard Deviation
0.12
0.42
0.72
0.61
0.71
0.12
-
-
0.23
Standard Deviation
0.18
0.36
0.47
0.32
0.34
0.09
-
- •
0.24
(Continued)
a ratio for nitrate set to 1.0, ammonium ratio set to 0.5
122
-------
Table 5-16. (Continued)
SBRP
S042'
Ca2+
Mg2+
Na+
K+
cr
°N03-
aNH4+
H+
Median
0.62
1.72
1.83
1.14
1.48
0.40
1.0
0.5
0.50
Mean
0.60
1.54
1.69
1.06
1.36
0.36
1.0
0.5
0.52
Standard Deviation
0.12
0.39
0.45
0.35
0.36
0.10
-
-
0.16
0 ratio for nitrate set to 1.0, ammonium ratio set to 0.5
123
-------
As reported by Church et al. (1989), preliminary NOAA estimates of wet and dry sulfur deposi-
tion for the NE (sites in central Pennsylvania, Whiteface Mountain, New York, and Rowland, Maine)
and the SBRP (Oak Ridge, Tennessee) were highly comparable to regional averages of the DDRP esti-
mates. The regional average of DDRP estimates of wet sulfur deposition in the NE was roughly 15
percent greater than the NOAA estimate for the sites examined and the DDRP estimate of dry sulfur
deposition was about 25 percent less than the NOAA estimate. The estimates of total sulfur deposition
in the NE were virtually identical; the total and even the individual component estimates (i.e., wet and
dry) in the SBRP were within 4 percent (B. Hicks, pers. comm.).
A comparison of regional (i.e., NE and SBRP) dryAotal deposition ratios as obtained from the
DDRP estimates and "as quantified by NOAA for the same region" showed remarkable agreement for
both regions for sulfate, nitrate, and ammonium (Church et al., 1989). No NOAA values were available
for chloride, we could make so no comparisons for that ion. The DDRP estimates of the dryAotal ratio
for base cations was generally just over twice as high as the NOAA values for the NE and ranged from
4 to 10 times as great in the SBRP. This difference is due at least partly to the fact that DDRP
estimates for base cation deposition included an estimate for coarse-particle dry deposition, whereas
the NOAA values did not. For example, a comparison of the average DDRP estimates of fine-particle
base cation deposition for five watershed sites in proximity to the NOAA West Point station showed
good agreement (i.e., within 20 percent for calcium, sodium, and magnesium; within 50 percent for
potassium) with the measured NOAA values at that site (B. Hicks , pers. comm.). To account in part
for the uncertainties associated with the DDRP estimates of base cation dry deposition, Church et al.
(1989) performed sensitivity analyses with datasets having much reduced base cation values. More
formal uncertainty analyses were performed with the integrated ^watershed models. In general, DDRP
analyses and the conclusions drawn from them were not sensitive to these uncertainties (see Sections
9 and 10, Church et al., 1989).
5.6.3.3 Sulfur Deposition Scenarios
Typical year total sulfur deposition (as sulfate) for the M-APP is shown in Plate 5-3. As
described in Section 5.6.1 (see Figure 5-12), the DDRP was requested to examine the effects of
scenarios of both current and altered sulfur deposition in the M-APP. The sulfur increases and
decreases were performed as sulfate with both dry and wet deposition altered at equal and constant
percentages (of the total) each year. No good existing data indicate any procedure other than
adjusting wet and dry H+ to coincide with changes in sulfur deposition (A. Olsen, R. Dennis, pers.
comm.). That was the procedure followed. Wet H+ was adjusted equal to the wet sulfate adjustment
and dry H+ was recomputed so that the sum of dry cation inputs was not less than the sum of dry
anion inputs (on an equivalent basis) in any month. Values of typical year sulfate deposition for all
DDRP study sites are shown in Plate 5-4.
124
-------
-------
Plate 5-3. Pattern of typical year sulfate deposition for the DDRP M-APP study sites.
125
-------
-------
APPALACHIAN REGION
Typical Year
Sulfate Deposition
-------
-------
Plate 5-4. Pattern of typical year sulfate deposition for all DDRP study sites.
126
-------
-------
DDRP STUDY REGIONS
Typical Year Sulfate Deposition
SOr2 (g/m2)
© < 2
©2-4
©4-6
® 6 - 8
• > 8
Mid-Appalachian
Region
Northeast
Southern Blue Ridge
Province
-------
-------
5.7 RUNOFF DATA
An estimate of average annual runoff for the DDRP study sites was necessary for all levels of
DDRP analyses. Given the site selection procedures used in the DDRP, it is not surprising that the
DDRP study sites are ungaged and that measured values of annual runoff are not available. Three
options existed for obtaining estimates of runoff. The first, gaging the systems, would not have been
practical, given the relatively short time frame of the DDRP and the large number of sites. The second
option was to use an interpolation method, such as kriging, to estimate runoff at each site. Large
variability in topography across the regions, and in other features that influence runoff, limited the
applicability of this method. Thus, we selected a third option for estimating runoff to each DDRP site:
interpolations were made from runoff contour maps developed from existing runoff data by expert
judgment of hydrologists experienced in runoff mapping.
5.7.1 Data Sources
Working in cooperation with the USGS, a runoff contour map of average annual runoff for 1951-
80 (Figure 5-13, Krug et al., 1990) was developed for interpolating runoff at the DDRP sites. The map
was developed to encompass the NE, M-APP, and SBRP regions of the eastern United States (Figure
5-13). Average annual runoff data for the 30-year period were taken primarily from watersheds of less
than 2,590 km2 having no diversions or regulations. If a gaging station did not have a complete set of
records for the 30-year period, then Krug et al. (1990) calculated a 30-year estimate using the stan-
dard correlation methods described by Matalas and Jacobs (1964). Runoff contours were plotted at a
scale of 1:500,000 at 5.1-cm (2-in) intervals up to 76.2 cm (30 in) and at 12.7-cm (5-in) contour inter-
vals for runoff greater than 76.2 cm (30 in). Krug et al. (1990) provide more specific information on
map development and quality assurance.
5.7.2 Runoff Interpolation Methods
A simple nearest-contour linear interpolation method was used to estimate runoff for each DDRP
site. The Krug et al. (1990) map was digitized into a GIS system by the USGS. By means of the GIS
(Campbell et al., 1989), the DDRP study sites were overlaid onto the runoff contour maps and runoff
was interpolated at each DDRP site to the nearest inch, based on the nearest contour to a site. The
nearest contour was determined using an engineer's scale to measure a line from the station location
perpendicular to the contour (Rochelle et al., 1989b).
5.7.3 Uncertainty Estimates
Determining a quantifiable estimate of the uncertainly associated with the runoff interpolations is
important to the effective use of the runoff data in the Levels I, II, and 111 analyses. Working with the
USGS, we conducted an analysis to estimate the uncertainties in using a runoff contour map to deter-
mine runoff at a specific site. This analysis was incorporated into the development of the 1951-80
runoff map (Rochelle et al., 1989b; Krug et al., 1990). We randomly selected a subset of the total
USGS sites avaijable for map development and withheld these sites from use by the USGS as they
developed the map. Then we used the runoff contour map to interpolate runoff at these sites and
127
-------
ANNUAL RUNOFF 1951 - 1980
Figure 5-13. Example of average annual runoff map for 1951-80 (Krug et al., 1990).
128
-------
compared the interpolated values to the actual long-term measured values. We determined that runoff
could be estimated, on the average, within approximately 8.9 cm (3.5 in. or 14.9 percent) of the actual
measured runoff. Rochelle et al. (1989b) provide a complete discussion of the uncertainty analysis.
We conducted a second analysis to test the consistency of interpolating runoff using the hand-
linear interpolation method described in Section 5.7.2. For the NE region, we plotted 883 NSWS
watersheds on the 1951-80 runoff contour map, and interpolated runoff to each site. Of the 883
NSWS sites, a subset of 146 watersheds was plotted onto the runoff contour maps, and runoff was
interpolated at the test sites a second time. We compared the two independent estimates to check for
consistency in using the hand-linear interpolation method. We found that 11 percent of the sites had a
2.5-cm (1-in.) difference between the two interpolations (5 percent runoff difference) and 1 percent had
a 5.1-cm (2-in.) difference between the two runoff interpolations. The results of a paired t-test indicate
that the hand interpolation method is reasonably consistent, with no significant differences in runoff
between the two iterations (t = 0.65, p = 0.51). Rochelle et al. (1989b) provide a description of these
uncertainty analyses.
129
-------
-------
SECTION 6
REGIONAL POPULATION ESTIMATION
6.1 INTRODUCTION
This section describes the procedures used to extrapolate analyses on individual watersheds to
the target populations in the study regions. This process of extrapolation is called population
estimation. •
6.2 PROCEDURE
6.2.1 Use of Variable Probability Samples
As was done for the DDRP sample in the Northeast (NE) and Southern Blue Ridge Province
(SBRP) (Church et al,, 1989), probability samples were selected for stream watersheds in the Mid-
Appalachian region (M-APP) (Section 5.2). Any quantity that can be defined for a sample unit (i.e., for
each watershed) can be extended to a corresponding population quantity through the probabilistic
structure of the sample. The quantity can be a measured variable or a model-based estimate. It can
be a number, a vector, or a function. In the National Stream Survey (NSS), most quantities were
measured values, and the measurement error tended to be small relative to the sampling variation. In
contrast to the NSS, many of the quantities produced in the DDRP are model outputs believed to have
significant uncertainty associated with them. The population estimation techniques provided in this
section apply to any probability sample with defined inclusion probabilities. Thus, they are applicable
to any identifiable subset of the DDRP sample. Explicit provision is made for including uncertainty
associated with the quantity that is extended to the regional population.
In the NSS, and hence the DDRP, the size of the target population is not precisely known. The
sampling frame for the NSS was an area/point frame based on USGS maps. There was not a popula-
tion of known size from which to draw samples. Thus, the size of the target population is estimated
from the sample size. This presents no particular difficulty, as long as each unit in the sample has a
known inclusion probability.
The design of the NSS and the DDRP also permits arbitrary subsetting of the sample. In some
cases, the subsetting may correspond to a redefinition of the target population (e.g., the exclusion of
streams affected by urbanization). In such cases, the inclusion probabilities for the remaining sample
units would not change, which implies a smaller target population. In other cases, the subset should
be viewed as a subsample. In these cases, a smaller sample would be used to make an inference
about the same target population, and the inclusion probabilities would change. This might occur if a
selected stream could not be sampled for some reason. Inferences could still be made about the
same target populations, but the inclusion probabilities would change.
130
-------
6.2.2 Estimation Procedures for Population Means
The structure of the DDRP sample is almost identical to the structure of the NSS sample. The
estimation procedures parallel those detailed in the Sampling and Analysis Plan for Streams in the
National Surface Water Survey (Overton, 1987). Let n be the size of the sample selected from the
target population, let p, be the probability that sample unit i was included in the sample, and let Py be
the joint inclusion probability of units i and j. For sample unit i, let y( be the "true" quantity, and let Zj
be the observed quantity, i.e., the unknown true value with an associated error, e-t. The error may be
an observation error or a measurement error; it could also be a prediction error. In each case, we
assume the characteristics of the error distribution are known, and that the uncertainty in the observed
values is characterized by that error distribution. The basic estimation procedures follow the Horvitz-
Thompson estimator (Cochran, 1977) for variable probability samples; some details, however, depend
on assumptions made about the observation error. Several distinct error models are treated here.
In one case, the uncertainty is due to an additive error term, so that the magnitude of the uncer-
tainty is constant over the range of the response. The observation is related to the true value through
the equation zl = Vj + e j. Two distributions were available to handle this case: the error term is
assumed to have either a normal distribution with mean 0 and variance a2 or a uniform distribution
over the interval (-a,a). For this uniform distribution, the mean is 0 and o2 = a2/3.
In a second case, the magnitude of the uncertainty depends on the magnitude of the response.
This can be modelled with a multiplicative error term, where the uncertainty is proportional to the
response, so that z, = y (e,. We assume that the uncertainty follows a log-normal distribution with a
mean value of 1 and a variance a2 = RSD2, where RSD is the relative standard deviation.
An implication of the above multiplicative model is that the uncertainty goes to 0 along with the
response. In some instances, however, there is appreciable uncertainty even when the response is 0.
For these cases, we assume that the uncertainty is proportional to the sum of the response plus an
offset (h), so that the observation equation is z, = yl + (y, + h)e s = ys (ej + 1) + hej. The mean value
of the error term is 0, and the o2 = RSD2. As above, we use a log-normal distribution for this case.
The error structure affects only the variance of the population total, the variance of the
population mean, and the estimator of the cumulative distribution function and its associated variance.
The estimator of the target population size and population total take the same form under all of the
above error structures.
* *
Estimator of population total, T:
T = Zz, /p.
(6-1)
Estimator of the size of the target population, M:
131
-------
N = 21/p,
1
(6-2)
Estimator of population average, ?:
Y = TIN.
(6-3)
Both T and N are random variables, and both are unbiased estimators of the respective popu-
lation quantities. However, ?, similar to most ratio estimators, is a slightly biased estimator of the
population average.
6.2.3 Estimators of Variance
For all three error models, the estimator of the variance of T has the form
n O-A
War (T ) = 2 _
z " "
' + 2 2
) zizi\
' '
p
/>/ Pi Pj Pif
(6-4)
where g(e,z) is a function that depends on the error model and the sample data. For the additive
model, g(e,z) = o2 N; for the multiplicative model, g(e,z) = o22zi2/pj, and for the multiplicative model
with offset, g(e,z) = a2 2(z, + h)2/pj, where h is the offset.
A
The variance of N is estimated by
77 (1 -Pi) '" "
Var (N ) = 2 ' + 2 2
p
/>/ PiPjPii
(6-5)
The joint inclusion probabilities, p^, are determined by the structure of the DDRP sample. They are
computed according to the algorithm in the Sampling and Analysis Plan for Streams in the National
Surface Water Survey (Overton, 1987).
Finally, the variance of the estimator of the population average is obtained from a first-order
variance propagation using Equations 6-4 and 6-5:
(6-6)
Var(Y) =
Var(A/)/AT - 2TCov(T,N)/N*,
where
132
-------
n n
Cov (f,N ) = 2 2 (p, P/ - P// )(1/P/ - 1/Py )(z;- /Pf - zy /P/2)
(6-7)
Confidence intervals (Cl) are derived from the usual normal theory, e.g., a 95 percent Cl on the
population average is given by
Y± 1.96\/Var (Y)
(6-8)
6.2.4 Estimator of Cumulative Distribution Function
Let N(y) be the total number jn the population with the value of Y < y, so that the
cumulative distribution function of Y is FY (y) = N(y)/N. An estimator of N(y) is
A/z(y) = 2 1/P/ = 2 v, (y) /Pi
(6-9)
where
v. (y) = {i,Zi^ y
' vy; X0, Zj > y
An estimator of the cumulative distribution function of Y is
Fv(y) = N(y)/N-
(6-10)
The variance of FY has both a sampling component and a component due to measurement uncer-
A A A
tainty. The variance of N(y) and covariance of N(y) and N are needed to calculate the sampling
variance of FY. These are given by
Var(N(Y))=FY(y)(l-FY(y)) 2 1/P;2 + FY2
(6-11)
and
Cov(N,N(y)) = FY(y)Var(N).
(6-12)
133
-------
Then a first-order variance propagation formula gives
Var(FY) = Var(N(y))/N 2 + N 2(y)Var(N)/N 4 - 2N(y)Cov(N(y),N)/N2
for the sampling variance.
(6-13)
A Monte Carlo procedure is used to calculate the measurement variance. The sampling vari-
ance and the measurement variance are added to obtain total variance. The median and quintiles of
A
the distribution of Y are estimated by the linear interpolation of FY.
6.2.5 Hypothesis Testing for Changes in Population Totals
In tables such as Tables 10-9 and 10-10, population totals are estimated using the methodology
outlined in Sections 6.2.2 - 6.2.4. It is natural to question whether the apparent changes in population
totals across deposition scenarios, or through time within individual deposition scenarios, are statistic-
ally significant. In each case, the question is whether the number of discrepant watersheds in the
table of population estimates is significantly larger (or smaller) than zero.
We have specific expectations about the direction of change between scenarios, based on our
understanding of the chemistry of the situation. In moving from a higher deposition scenario to a
lower deposition scenario, we expect to see either minimal change in surface water chemistry or an
increase in surface water ANC. In this case, the difference of the population totals would be expected
to remain constant or decrease. For this reason, we may treat all rejection regions for the hypothesis
tests as one-sided.
To estimate the number of systems in the target population showing changes between two
deposition scenarios, we can look at the difference between the projected model values for the two
scenarios at a particular site. Because of the aforementioned scientific considerations, we know in
theory which way to compute the difference so that the difference will be 1 if the model values indicate
a change between scenarios, and 0 otherwise. Then the total number of systems in the target
population can be estimated by using the Horvitz-Thompson estimator:
d,/p,
(6-14)
where dj can be written as
dj = I{ANC1 < 0} - I{ANC2 < 0},
with l{ }, the characteristic function, taking the value 1 if the quantity inside the brackets is true, and 0
otherwise.
134
-------
Similarly, to estimate the number of systems in the target population whose chemistry would
change over time for a particular deposition scenario, we can examine the difference between the
characteristic functions of the projected model values of ANC for the scenario at two different times.
Here also, because of the scientific considerations, we know in general which way to compute the
difference so that the difference will be 1 if the model values indicate a change over time, and 0
otherwise. Again, the total number of systems in the target population can be estimated by using the
Horvitz-Thompson estimator.
A
Now we must estimate the variance of T. This variance should include not only the standard
terms in the variance estimates for variable probability sampling, using the Horvitz-Thompson
estimator of the variance expressed previously in this section, but also an estimate of the variance
component that is due to the variability of the model inputs and the model sensitivity to these inputs.
We obtained this second component from the individual model runs performed during the cali-
bration of the Model of Acidification of Groundwater in Catchments (MAGIC). For each watershed,
there were up to 10 acceptably calibrated runs, and a watershed was considered uncalibrated unless
there were at least 5 such runs. The final watershed value given by the model was the median of the
values for these runs.
The program (i.e., MAGIC) used a Latin Hypercube sampling algorithm, which is a stratified
Monte Carlo sampling method, to select the vectors of values for the model runs. Variance estimates
of the inputs were incorporated in selection of the vectors. For this reason, we decided that it was
acceptable to treat the variance of the model outputs as an estimate of the within-watershed variance
due to input variability and model sensitivity.
The analysis was performed on the differences of the characteristic functions rather than on the
original ANC. Therefore, we computed this variance on the differences of the characteristic functions
in the individual model runs. For comparisons between scenarios, we matched the model runs across
scenarios such that the input variables other than deposition were matched (J. Cosby, pers. comm.).
For comparisons over time, we took the difference between the values of the characteristic function in
the same model run at the two years of interest.
Because we worked with differences of characteristic functions within watersheds, we treated the
data as independent observations on a (possibly degenerate) standard trinomial distribution, that is, a
discrete distribution taking only the values 1, 0, and -1. The calculated variance is that of one obser-
vation, not the median of 5-10 observations, some of which may have been excluded as part of unac-
ceptable runs. We did not have an estimate of the effect of this censoring, and therefore we have not
calculated the variance of the median of censored observations from a trinomial distribution. Instead,
we used the calculated variance as an overestimate of this component of the variability. The effect on
the overall variance appeared to be negligible, except in those cases where the watershed differences,
dj, were identically zero. In all these cases, the change in population totals is zero, and hence not a
statistically significant change, regardless of the magnitude of this component of the variance.
135
-------
We assumed an additive model for the variance estimates. Because we worked with the d|
rather than the ANC, we did not have problems with skewed data, and a multiplicative model was not
necessary. Therefore, for the term representing the total variance due to input variability and model
sensitivity, we used s2 N, where s2 is the pooled variance estimate obtained by taking a weighted
average of these within-watershed variances, with the inverses of the watershed inclusion probabilities
used as the weights.
A
We used the Horvitz-Thompson estimator of the variance of T, since we were working with
subsets of the regions. We used the py estimates discussed in Section 6.2.3.
The sum of the Horvitz-Thompson variance estimate and the model sensitivity term provides an
estimate of the total variance in the estimate of the difference in the target population. The square root
of this sum was treated as the standard deviation, s-}-, of T, and a standard z-test
z = (T - o)/sf.
(6-15)
was used as an approximation. As stated earlier, we have scientific reasons to expect change in a
particular direction, so a one-sided hypothesis was used in each case. The usual p-values for one-
sided normal tests were generated in SAS as part of the calculations.
6.3 UNCERTAINTY ESTIMATES
The quantities displayed in this report are the end result of a sequence of operations, beginning
with the collection of a physical sample in the field and ending with the production of a table or graph.
We conducted chemical analyses, data aggregation, data reduction, and processing of the data
through various mathematical models. The final result contains an element of uncertainty that
originates, first, in the design, then in the implementation of the field protocol, and finally in the
precision of the basic measurement process (e.g., the chemical analytic precision). The uncertainty
on the final result can be quantified by propagating the uncertainty, or its mathematical analog,
through the same sequence of operations as the data.
In the DDRP, we have used several techniques to propagate uncertainty through a functional
relationship, which could be a complex simulation model as well as an explicit function. Let f(x1( x2
xn) be a function of the variables x1 ,x2 xn, with uncertainties e.,, e2, ..., en, respectively. The
probability distributions, or at the least the variances, of the uncertainties are presumed known. If the
functional relationship is such that partial derivatives can be easily obtained, then the variance of
functional values can be estimated using a first-order linear approximation to the functional
relationship:
Var(f) =
(6-16)
In the case of a simulation model, the function is the model itself, and the partial derivatives
cannot be calculated explicitly. An approximation to the partials can be obtained by perturbing the
in turn. If a suitably small perturbation is chosen, then the ratio of the change in output to the
136
-------
perturbation is an estimate of the partial derivative. These estimates can then be used in a first-order
propagation, as above.
A disadvantage of both these techniques is that they ignore possible correlations among the
uncertainties. One way to account for such correlations is to propagate not only variances but also
covariance terms. The "first-order, second-moment" technique used in the Enhanced Trickle Down
uncertainty analysis is a means of doing exactly that. A first-order approximation is made to the
model, and Kalrnan filtering techniques are used to build up an estimate of the state variable variance-
covariance matrix. A final method that we used in uncertainty assessment was Monte Carlo. The
Monte Carlo method is applied by repeatedly calculating the value of f, each time perturbing the value
of each Xj by a random quantity drawn from the respective uncertainty distribution. Monte Carlo is
most easily appSied when uncertainties are statistically independent, but it can also be applied when
correlations exist. A variant of Monte Carlo, called "fuzzy optimization", was used in the uncertainty
analyses for MAGIC.
6.4 APPLICABILITY
This section discusses the Level II and III population estimation approaches for DDRP, including
the statistical formulas used to estimate population means, variances, and cumulative frequency
distributions. The population estimation procedures are generic and do not depend on the level of
analysis. The specific target populations for inference, however, do depend on the analyses
performed. Not all DDRP watersheds were used at each level of analysis, so the target population
varies. The explicit target populations being considered in the analyses are discussed in Sections 9
and 10. The generic uncertainty estimation procedures introduced in this section also are more
explicitly discussed for each of the individual analyses in Sections 9 and 10.
137
-------
SECTION 7
WATERSHED SULFUR RETENTION
7.1 INTRODUCTION
The fate of sulfur deposited in a watershed is important in determining the response of the
associated surface water because sulfate can act as a mobile anion in the soil matrix (see Section 3).
In systems at steady state with regard to sulfur deposition (i.e., inputs = outputs or zero net retention),
the leaching rate of either basic or acidic cations by the "carriefanion" sulfate has been maximized.
Given no increase in sulfur deposition, future acidification (loss of ANC) of these systems would be
determined principally by cation leaching and the possible depletion of the soil exchange complex. In
systems below sulfur steady state (i.e., inputs > outputs), the acidifying effect of sulfate-driven cation
leaching has not been maximized. As sulfate leaching increases in these systems, soil adsorption
sites are filled on a net basis, and acidification and the rate of acidification increase over time. A
circumneutral lake or stream draining a watershed with positive net sulfur retention will continue to
acidify, possibly becoming acidic (i.e., ANC < 0), as long as rates of sulfur deposition (inputs) exceed
outputs. Thus, even if sulfur deposition decreases, some circumneutral systems will acidify and may
become acidic. Knowing the patterns of watershed sulfur retention, therefore, is important with regard
to understanding and forecasting the potential effects of sulfur deposition on surface water chemistry.
In this section, we examine regional patterns of sulfur retention, as estimated using input/output
budget analyses.
The purpose of the watershed sulfur retention component of the DDRP Level I analyses is to
estimate the current status of annual sulfur retention in watersheds of the eastern United States, with
primary emphasis on the NE, M-APP, and SBRP regions. The M-APP Region provides important infor-
mation for the interpretation of sulfur retention patterns for the NE and SBRP, as well as a more
complete regional perspective for upland areas of the entire eastern United States. Specific objectives
of these analyses are to:
• characterize current average annual input/output budgets in the NE, M-APP, and SBRP using
(1) data from intensively studied sites and (2) estimates computed from regionally extensive
datasets; and
• compare annual sulfur retention patterns within and among regions to determine possible
trends relative to water chemistry, soils, and atmospheric deposition.
Many of the analyses presented here were reported with DDRP results for the NE and SBRP region
(Church et al., 1989), but are repeated here for completeness and to provide a context for discussions
of sulfur dynamics in subsequent sections (9.2 and 10). The previous DDRP report also provided (1) a
detailed assessment of sulfate reduction in wetlands and in-lake sediments as a watershed sink for
sulfur, (2) a detailed assessment of internal sulfur sources in the DDRP watersheds, and (3) a region-
alized uncertainty analysis for sulfur input-output budgets. These issues are summarized briefly in this
section; readers are referred to Church et al. (1989) for a more complete discussion of those topics.
138
-------
7.2 RETENTION IN LAKES AND WETLANDS
Church et al. (1989, Section 3) identified and discussed several processes that can cause sulfur
to be retained within watersheds. One of the processes considered was retention by sulfate reduction
in wetlands and/or lakes. Retention in these environments occurs principally by dissimilatory reduction,
with sulfate used as an electron acceptor and with hydrogen sulfide, organic sulfur, or metal sulfides
as end products (Rudd et al., 1986a,b; Brezonik et al., 1987). Analyses in the DDRP (Shaffer et al.,
1988; Shaffer and Church, 1989; Church et al., 1989) were designed to estimate in-lake sulfur retention
in the ELS populations of drainage lakes and impoundments; in the NE, SBRP, and Upper Midwest
(UMW) (Unthurst et al., 1986; Kanciruk et al., 1986), retention in seepage and closed lake basins was
not evaluated due to high uncertainty in the hydrologic budgets for those systems. ELS sulfate data,
along with precipitation and runoff estimates generated within the DDRP (or by ELS for the Upper
Midwest), were used with an in-lake sulfur retention model developed by Baker et al. (1986b) and Kelly
et al. (1987) to estimate percent sulfur retention. Results showed that in all three regions, median in-
lake sulfur retention was 5 percent or less; in-lake retention was estimated to exceed 10 percent of
deposition in only 13 percent of NE lakes and in fewer than 1 percent of lakes (mostly impoundments)
in the Southern Blue Ridge.
Regression analyses were used in the DDRP.to evaluate sulfur retention in wetlands and wet
soils for the NE; results indicated a statistically significant positive correlation between the proportion
of wetlands on a watershed and the percent of watershed sulfur retention. For many reasons, quanti-
tative use of these results or of any similar generalization of wetland area-sulfur budget relationships is
difficult, and we have not attempted any such use. The importance of wetland retention on watershed
sulfur budgets depends on the location of the wetland in the watershed and the portion of watershed
runoff flowing through it. Also, sulfur reactions in wetlands and wet soils can change seasonally or in
wet/dry years. Wetlands and wet soils can act as sulfur sinks (reduction of sulfur) during wet periods
when the system is anaerobic, but can become major sulfate sources due to reoxidation of sulfides
upon drying (Bayley et al., 1986; Nyborg, 1978). Interpretation of these analyses was further compli-
cated by high uncertainties in both the independent (% wetlands) and dependent (% S retention) vari-
ables in the regression equation.
Qualitative evaluation of soil and hydrologic characteristics on DDRP M-APP watersheds sug-
gests that sulfate reduction will not be an important long-term sink on most watersheds. Histosols
occur on only five watersheds in the M-APP sample, and represent less than 2 percent of watershed
area on any of those catchments. Poorly drained soils, subject to seasonal flooding, cover more than
5 percent (maximum = 19 percent) of the area on 15 watersheds in the M-APP sample. These soils
could provide reducing environments for sulfate reduction during wet periods, but any sulfate seques-
tered during those times would in all likelihood be quickly reoxidized upon drying and reaeration of the
soil. Based on these low proportions of permanent anaerobic zones on the watersheds, the potential
for net sulfate reduction in M-APP watersheds appears to be small relative to other sulfur sources/
sinks, and we do not address it further here.
139
-------
7.3 WATERSHED SULFUR RETENTION
Two approaches have been used in the DDRP to characterize regional patterns of watershed
sulfur retention. The first approach was to compile and review sulfur inputyoutput budgets at
intensively studied sites (Rochelle et al., 1987). Figure 7-1 summarizes the findings of this review.
Definitive statements about region-wide sulfur retention could not be made on the basis of these data,
however, because of lack of spatial coverage by the intensively studied sites and inconsistencies in
data used for budget calculations. There are, however, clear trends in sulfur retention from north to
south in the eastern United States, especially relative to the extent of the Wisconsinan glaciation, with
higher retention in the southern areas (Figure 7-1). The DDRP Level I sulfur retention analysis
examines these apparent trends in more detail using a second approach, which has involved gener-
ating regionally consistent sulfur input and output data (Section 5) for the surface water sites sampled
by the Eastern Lake Survey (ELS) and National Stream Survey (NSS), then characterizing the sulfur
budget status for the representative NSWS populations of watersheds on a region-wide basis.
7.3.1 Methods
7.3.1.1 Input-Output Calculation
Level I sulfur retention analyses use an annualized mass balance approach to estimate percent
retention. The general equation used to calculate percent sulfur retention is:
% Retention = ((Sw + Sd - R * SS)/(SW + Sd ))*100 (7-1)
where: Inputs
Sw = wet sulfur deposition (mass/length2/time)
Sd = dry sulfur deposition (mass/length2Aime)
Outputs
R = runoff (length/time)
Ss = surface water sulfur (mass/length3)
Equation 7-1 relates the total sulfur input (on a mass basis) to total sulfur output for each watershed.
We applied this equation to all ELS lake systems (except seepage and closed lakes) and NSS water-
sheds in the regions of interest.
7.3.1.2 Data Sources
7.3.1.2.1 Inputs-
Wet sulfur deposition was estimated for each site using chemistry data from the National Trends
Network/National Acid Deposition Program (NTN/NADP) network and precipitation data from the
National Oceanic and Atmospheric Administration (NOAA), National Climatic Data Center (NCDC)
network (Section 5.6). Briefly, wet sulfate concentrations and precipitation were kriged to each site,
and wet deposition was calculated (see Wampler and Olsen, 1987, for a detailed description of the
140
-------
\\» 1 '-'"r~"
SULFUR RETENTION
RETENTION
RETENTION
RETENTION
Figure 7-1. Percent sulfur retention for intensively studied sites in the United States and
Canada relative to the southern extent of the Wisconsinan glaciation (adapted
from Rochelle et al. (1987).
141
-------
calculation). Dry sulfur deposition was estimated based on output from the Regional Acid Deposition
Model (RADM) (see Section 5.6). Several deposition data sets have been developed during the DDRP
to meet a variety of modelling needs; for a thorough description of the nature of each data set and the
differences among them, see Section 5.6 of Church et al. (1989).
7.3.1.2.2 Outputs-
We used estimates of long-term average annual runoff based on data for the 30-year period of
1951-80 (see Section 5.7 for details). For these analyses we used NSWS sulfate data, which measure
only dissolved inorganic sulfate; it represents the vast majority of sulfate in most surface waters and
does not significantly compromise analyses (David and Mitchell, 1985; Mitchell et al., 1986). Section
5.3 discusses the surface water chemistry data used in these analyses. For additional information
concerning the ELS and NSS surface water sulfate estimates, see Linthurst et al. (1986), Messer et al.
(1986), and Kaufmann et al. (1988). Seepage lakes and closed lakes were excluded from the
analyses.
7.3.2 Uncertainty Estimates for Sulfur Input-Output Budgets
As part of sulfur input-output budget analyses, Church et al. (1989) estimated the uncertainty
associated with calculations of average annual sulfur retention using Monte Carlo analysis, similar to
that described in Section 6.3. Uncertainties for individual input/output variables in Equation 7-1 were
determined, then propagated through the retention equation to determine an estimate of overall uncer-
tainty of the percent retention calculations. A total of 10,000 iterations of the budget calculation were
made for each of 34 randomly selected watersheds in the NE, M-APP, and SBRP. From those results,
Church et al. (1989) developed a multiplicative normal relationship between steady-state sulfate and
uncertainly, which can be used to identify catchments for which sulfur retention or release is sig-
nificantly different from zero:
Est. std. dev. = - 4.9 + 0.339
2-
JSS
(7-2)
Results of this analysis were very useful for identifying watersheds with putative internal sources of
sulfur. We excluded these watersheds from subsequent analyses. An alternate approach for identi-
fying internal sources of sulfur, which focused on bedrock mineralogy, was largely unsuccessful due
to the spatial resolution of state-scale geologic maps. For example, the location of unit boundaries
was uncertain, and small sulfide-bearing bodies either were included in a general map unit or not
delineated at all. Watersheds having internal sources of sulfur were ultimately identified on the basis
of (1) the uncertainty estimates described here, (2) identification of outliers/influence points in
regression analyses, and/or (3) mapped delineation as strip mines, borrow pits, etc. For the two
regions (i.e., M-APP and SBRP) having a majority of watersheds with significant sulfur retention, we
focus subsequent analyses on assessing the importance of adsorption as a retention mechanism
(Sections 9 and 10). For a thorough discussion of uncertainty analyses and of identification of water-
sheds suspected of having internal sources of sulfur (including a list of watersheds dropped from
budget analyses), see Church et al. (1989, Sections 7.3.2 and 7.3.3).
142
-------
7.3.3 Input-Output Budgets - Results and Discussion
This section summarizes, at a regional scale, the calculated percent sulfur retention for sites
located in the NE, M-APP, SBRP, and several adjacent regions. We eliminated from the analyses sites
identified as having internal sulfur sources through the steady-state sulfate concentration analysis
(listed in Table 7-6 of Church et al., 1989). In addition, three sites in the SBRP that were sampled as
part of the Pilot Stream Survey were dropped due to outlier surface water chemistry (i.e., ANC > 1000
/90 percent of watershed area) was covered by wetlands or chronically flooded
soils. Results for the region indicate that most watersheds in the NE are at or near steady state with
respect to atmospheric deposition of sulfur. These analyses also indicate, however, that at a few sites
in the region, sulfate reduction in soils or lake sediments acts to maintain sulfate concentrations at a
level below steady state.
7.3.3.2 Mid-Appalachian Region
The Mid-Appalachian Region does not present as clear a picture of percent sulfur retention as
the NE (Table 7-1; Figure 7-2B). Retention for individual watersheds varies widely within subregions,
exceeding 90 percent in several cases, while net release is > 80 percent in catchments in several
subregions. For this study, we have defined the M-APP Region as a combination of NSS subregions
2Bn and 2Cn (Plate 7-1). Kaufmann et al. (1988) defined these regions as the Valley and Ridge and
Northern Appalachians, respectively. We found that percent retention was widely distributed within
each subregion, with no clear intraregional patterns of low or high percent net retention. In general,
for both subregions, net retention is low, averaging less than 30 percent. On average, retention in
143
-------
Table 7-1. Percent Sulfur Retention - Summary Statistics by Region for Lakes and Stream
Reaches in the NSWS (ELS) and NSS
Region
NE
ELS Rg. 1b
1A
1B
1C
1D
1E
Mid-App
NSS 2Bn
NSS 2Cn
SBRP
PSS
ELS Rg. 3a
Misc.0
S. App. Pla,
NSS2X
Piedmont
NSS3A
Mid Atlantic Coastal
NSS SB
Poconos/Catskills
NSS 1D
Na
5,828
1,099
1,285
1,190
966
1,288
12,580
6,478
2,031
247
4,329
7,199
Plain
9,535
2,724
Mean
-5
-12
9
-4
-12
-9
28
-4
68
68
43
68
31
-22
Median
-5
-14
8
-7
-9
-12
40
3
75
79
50
78
34
-29
Std. Dev.
28
23
26
27
30
27
43
32
23
33
38
24
38
31
Min.
-70
-64
-66
-66
-70
-62
-83
-83
— *54.
\JT
-64
-64
-10
-60
-71
Max.
73
61
73
63
54
51
91
55
88
93
86
92
93
67
a Estimated target population calculated using NSWS weights (see Linthurst et al., 1986a; Messer et al., 1986a;
Kaufmann et al., 1988 for information on weights).
b Lakes sampled by the Eastern Lake Survey (ELS). All other sites represented in this table are stream reaches.
c Additional regions sampled as part of NSS Phase I (see Plate 7-1).
144
-------
NE
1.0r
O 0.8
O
0.
S 0.6
«= 0.4
as
1
o o-2
Upper Bound
Projected
Lower Bound
0.0
-100 -50 0 50
Percent Suffur Retention
100
Figure 7-2. F'opulation-weighted distribution of projected percent sulfur retention, with
upper and lower bounds for 90 percent confidence intervals, for NSWS target
population of (A) lakes in the Northeast, and for stream reaches in (B) the
Mid-Appalachian, and (C) Southern Blue Ridge Province Regions.
145
-------
Plate 7-1. Sulfur retention and wet sulfate deposition for National Surface Water Survey
subregions in the eastern United States.
146
-------
-------
NSWS SUBREGIONS
yEDIAN PERCENT SULFUR RETENTION
AND WET SULFATE DEPOSITION
2.25
MEDIAN PERCENT
SULFUR RETENTION
o - 20
20 - 40
40 - 60
60 - 80
80 - 100
Average Annual
Wet Sulfaie ^ 2.75-
Deposliion (g nf2 yr"1)* 3.06-
3.25
Eosiern Loke Survey
-2.25
Subregion X Retention
1A
16
1C
ID
IE
-14
8
-a
-12
2.00
Notional Stream Survey
liedian
Subregion X Retention
2Cn
26(1
3B
n
Us
3A
3
40
34
50
75
78
*DeP05i-t:ion for 1980 - 1984
(A. Qlsen. Personal Communication)
-------
-------
subregion 2Cn is estimated to be lower than in subregion 2Bn. Subregion 2Cn receives higher sulfur
deposition than does subregion 2Bn (Plate 7-1). Subregion 2Cn also has a high incidence of systems
with potential acid mine drainage influence, but systems identified by Kaufmann et al. (1988) or
identified by steady state analyses as haying probable internal sulfur sources (Church et al., 1989,
Section 7.3.3) have been dropped from this analysis and presentation of results, and we do not
believe that they significantly affect our results.
The Southern Appalachian Plateau and the Mid-Atlantic Coastal Plain have percent retention on
the average of 30 to 40 percent (Table 7-1; Figure 7-3). Although these regions show a large amount
of scatter in percent sulfur retention, they are characterized as having a majority of systems with
substantial net sulfur retention. Retention in these areas is also consistent with the general trend of
increasing sulfur retention along a north-to-south gradient. The geographic retention gradient is, in
turn, attributable to differences in soil age and physico-chemical characteristics that result in increased
sulfate adsorption along a comparable north-south trend, as is discussed in Section 9.2.
7.3.3.3 Southern Blue Ridge Province
Median net sulfur retention for the SBRP is approximately 75 percent (Table 7-1; Figure 7-2C).
Rochelle and Church (1987), working with sulfur deposition data from Water Year 1984, found similar
results. The average percent sulfur retention for the Piedmont Region (adjacent to the SBRP) is also
high compared to the NE and M-APP (median = 78.0, Table 7-1 and Figure 7-3).
7.3.4 Regional Patterns of Sulfate Concentration and Steady-State Sulfate
Comparison of sulfur budgets for the three major DDRP regions in the eastern United States
shows clear regional differences in budget status. Budgets in the NE are, on average, very close
to sulfur steady state; there is high net retention in the SBRP. Retention is highly variable in the
M-APP Region, ranging from high retention in some systems to high net release in others. Viewed
differently, budget data indicate that the relative potential for increases in surface water sulfate
concentration as watersheds come to steady state (assuming current deposition) is small in the NE,
intermediate (for the region) and highly variable (for individual watersheds) in the M-APP, and high in
the SBRP.
In considering the potential for future effects of acidic deposition on soils or surface waters, it is
important to consider not only budget status (i.e., the potential for relative changes in sulfate), but also
the potential for absolute change in sulfate (i.e., the -difference between present and steady-state
concentrations). Figure 7-4 and Table 7-2 present data for DDRP sample watersheds for sulfate
concentration, steady-state sulfate concentration, and percent sulfur retention. The sulfur distributions
shown here are somewhat different from those presented earlier in this section because the previous
data describe regional ELS or NSS lake/stream populations, whereas these data represent smaller
DDRP target populations, from which we have excluded certain subsets of lakes/streams (e.g.,
seepage lakes, systems with high ANC, watersheds larger than 30 km2).
147
-------
Southern Appalachian Plateau A
Mid-Atlantic Coastal Plain
B
to
o
0.8
o
0.
o
fc
a
E
o
0.0
-100
Upper Bound
Projected
Lower Bound
•so o so
Percent Sulfur Retention
100
to
O 0.8
3=
O
a.
-------
LAKE/STREAM SULFATE
STEADY STATE SULFATE
1.0
100 150 200 250
SULFATE ftieq/L)
300
350
o
0.0
50 100 150 200 250 300 350 400 450
SULFATE (neq/L)
PERCENT SULFUR RETENTION
CHANGE IN SULFATE TO
STEADY STATE
1.0
o
0.0
-60
-i r
-40 -20 0 20 40 60 80
PERCENT SULFUR RETENTION
100
O
I— 0.8-
CC
2
O 0.6 i
CL
111
> 0.4-
0.2-
O
0.0
-50 0 50 100 150 200
DELTA SULFATE (neq/L)
250
Figure 7-4. Comparisons of regional distributions of (A) measured sulfate concentration,
(B) steady-state sulfate concentration, (C) percent sulfur retention, and (D)
change in sulfate concentration to steady state for DDRP target populations of
lakes or streams in the NE, M-APP, and SBRP.
149
-------
Table 7-2. Summary Statistics for Distributions of Sulfate Concentration, Steady-State Sulfate
Concentration, Change in Sulfate Concentration to Steady State, Percent Sulfur
Retention for DDRP Target Population Watersheds in the NE, M-APP, and SBRP
Variable
SO4 concentration
(Meq/L)
Steady-state SO4
(«eq/L)
Change in sulfate
to steady state
(Keq/L)
Percent sulfur
retention
Region
M-APP
NE
SBRP
M-APP
NE
SBRP
M-APP
NE
SBRP
M-APP
NE
SBRP
Mean
148
110
37
246
118
122
98
8
85
37
4
71
Std.Dev
55
40
26
72
47
23
-
-
24
20
17
Min.
55
34
15
140
50
86
-39
-51
25
-21
-54
17
Median
153
105
24
215
111
120
107
-0
83
44
0
78
Max.
313
214
119
405
239
203
248
28
167
74
69
87
150
-------
Steady-state sulfate concentrations in the NE and SBRP are similar, with medians of 111 and
120 ^eq/L, respectively. In contrast, due to the higher sulfur deposition to M-APP watersheds, steady-
state sulfate concentrations (median = 215/weq/L, maximum = 405/
-------
lowsulfate adsorption capacities (Section 9.2; see also Church et al., 1989; Rochelle et al., 1987).
Soils of the M-APP Region are predominantly Inceptisols and Ultisols. Measured sulfate adsorption
characteristics of soils from the three regions, discussed in Section 9.2, clearly demonstrate regional
differences in adsorbed sulfate pools and potential adsorption. Given the current patterns of wet
sulfur deposition (Plate 7-1) and assuming that the M-APP Region has received elevated levels of
sulfur deposition for a considerable period of time (Section 9.2), it appears that sulfate breakthrough
from the soils has occurred and that the region is in transition to lowered percent net sulfur retention
and significantly elevated surface water sulfate concentrations. There is little doubt that the increases
in sulfate are a direct consequence of anthropogenically derived increases in atmospheric sulfur
deposition.
The onset of high rates of sulfur deposition occurred later in the SBRP than in other regions
(Gschwandtner et al., 1985). That late onset, coupled with the high sulfate adsorption capacity of soils
in SBRP watersheds, has allowed the SBRP to experience a "lag" or "delay" phase in response to
acidic deposition. Model projections for the SBRP and sulfate trend data for monitored watersheds in
the region indicate, however, that soils and streams in the SBRP are beginning a transition to higher
sulfate outputs as sulfate begins to break through the soil column. The dynamics of sulfate transitions
in the NE and SBRP were described by Church et al. (1989) (Sections 9 and 10); we present summa-
ries of model projections for those regions, along with a detailed discussion of projections for the
M-APP Region, in Sections 9 and 10 of this report.
152
-------
SECTION 8
LEVEL I STATISTICAL ANALYSES
8.1 INTRODUCTION
The chemistry of surface waters in natural systems results from the integrated effects and
interactions of inputs from deposition, terrestrial processes, and in-lake or in-stream processes. In the
design of the DDRP, a sequence of four project objectives and three types of data analyses and model
applications were defined to characterize and quantify linkages between deposition, watershed
processes, and surface water chemistry (see Section 4). The first type of analysis (Level I) was
designed as a set of statistical analyses, principally multiple regression, that would facilitate
identification and understanding of associations between soil and landscape attributes and current
surface water chemistry. Level I analyses address the first two DDRP objectives: (1) to provide a
regional description of soil and watershed characteristics, and (ty to characterize relationships
between watershed attributes (soil chemistry, geology, vegetation, etc., along with deposition) and
surface water chemistry. Because much of our understanding of watershed-surface water chemistry
relationships has been based on site-specific data, which might or might not be representative of
regional conditions, an important objective of Level I statistical analyses in the DDRP was either to
corroborate, at regional scales, those fundamental understandings and assumptions (i.e., associations
between watershed attributes and surface water chemistry) or to identify additional important rela-
tionships not presently incorporated in conceptual or mathematical watershed chemistry models used
for Level II and 111 analyses.
In previous DDRP analyses for the NE and SBRP (Church et al., 1989), we conducted an exten-
sive Level I assessment of relationships between soil chemistry and/or groups of landscape variables
(geomorphic and hydrologic variables, depth to bedrock, land use and vegetation, bedrock geology,
and areal extent of soil sampling classes) and surface water chemistry. We evaluated landscape
attributes both as individual data sets and as an integrated set of variables. Results of those analyses
are summarized in this section. We did not conduct an analogous set of analyses for the Mid-
Appalachian (M-APP) Region. This decision was based on two principal considerations. First, the
analyses were extremely time consuming. Second and more important, there were no real surprises
in the results of the Level I analyses for either the NE or the SBRP. The results generally confirmed
the importance, at regional scale, of previously established relationships between surface water
chemical constituents and watershed attributes. A detailed description of Level I regression analyses
and results for the NE and SBRP is presented by Church et al. (1989), Section 8.
8.2 SUMMARY OF NE AND SBRP APPROACH AND RESULTS
The general approach used for Level I analyses was multiple regression, in most cases con-
ducted using stepwise procedures (SAS Institute, Inc., 1987, 1988). Residual analysis allowed us to
identify and remove outliers and leverage points from regression analyses. Five surface water chemis-
try variables (sulfate concentration, percent sulfur retention, ANC, pH, and sum of base cations [sum
of Ca + Mg in some cases]) were used as dependent variables in the regression analyses. Five
different groupings of independent variables were used:
153
-------
• Deposition to the watershed.
• Deposition plus one set of landscape attributes. Five'sets of landscape attributes were
developed and used, including: (1) geomorphic/hydrologic variables, (2) land use/vegeta-
tion, (3) depth to bedrock, (4) bedrock geology, and (5) soil sampling classes. Several
approaches were used to quantify individual attributes: (1) areal extent of each member of a
class of attributes (e.g., percent of watershed area with a particular bedrock type,
percentage in each depth-to-bedrock class) used for bedrock geology, sample class, and
depth to bedrock, (2) measured values, average value, or critical value (e.g., minimum or
maximum) used for geomorphic variables, derived hydrologic variables, bedrock sensitivity
class, and depth to bedrock, and (3) a set of factors developed from principal component
analysis (land use/vegetation).
• Soil chemical and physical properties, using values aggregated to watershed scale.
• Integrated analysis using deposition and all landscape attributes.
• Integrated analysis, using deposition, soil chemistry, and all landscape attributes except
extent of soil sample classes.
The results of Level I analyses for the NE and SBRP were generally consistent with the results of
similar regression analyses, and with our conceptual understandings of watershed attributes (and
implicit processes) influencing surface water chemistry (e.g., Hunsaker et al., 1986; Church and
Turner, 1986; Jenne et al., 1989). Deposition, soil physical and chemical properties, watershed
physical attributes, and anthropogenic factors were all correlated with one or more major surface water
chemistry variables. Table 8-1 summarizes results for a small subset of the analyses for the NE and
SBRP, namely those groups of attributes thought to most strongly influence surface water chemistry
and/or that exhibited the highest correlations. At regional scales, other groups of attributes (i.e.,
geomorphic and hydrologic variables, bedrock geology, and depth to bedrock) were less effective
than the attributes included in Table 8-1 in accounting for variability in surface water chemistry. For
example, the R2 values found in analyses using bedrock geology were not high, even though bedrock
is the ultimate source of base cations and alkalinity in watersheds, and is widely regarded as one of
the most important factors affecting watershed sensitivity to acidic deposition. This result is not
surprising in light of the large measurement uncertainties in the bedrock lithology data. The poor
observed relationship with geology is a result, presumably, of several factors: (1) only surface waters
with low ANC were included in the analyses, so the range of values for the dependent variables most
affected by weathering rates (pH, ANC, and sum of base cations) is relatively small, (2) state-scale
geologic maps were used, thus uncertainty in the position of map unit boundaries was high and
detailed information on spatially heterogeneous lithologies was not reliable at watershed scale.
Similarly, values of R2 for other analyses using groups of attributes with large measurement error, such
as atmospheric deposition, are also likely to be attenuated.
Sulfate concentrations in the NE most clearly show the direct effects of acidic deposition on
surface water chemistry. Net annual retention of atmospherically deposited sulfur in most NE
watersheds is low and surface water sulfate concentrations for lake systems in the DDRP sample (i.e.,
with ANC < 400 /teq/L) are highly correlated with rates of sulfur deposition. Sulfur deposition is the
most important predictive variable for surface water sulfate concentrations in watersheds in the DDRP
154
-------
Table 8-1. Summary of Level I Regression Analyses for DDRP Watersheds in the NE and
SBRP, Listing Values of R2 for Analyses Using Different Groups of Watershed
Attributes8
Attributes in
Region Regression
NE Deposition
Depn., land use
Depn., soil sample
class0
Dependent Variable
[S04=]
0.38
0.50
0.57 (141)
Percent S
Retention
nsb
0.19
0.41 (140)
ANC
0.16 (145)
0.37
0.51 (142)
PH
ns
0.30
0.46
SBase Cations
ns
0.42
(142) 0.52
Depn., all mapped
attributes'1
Soil chemistry
Depn., soil chemistry,
mapped attributes
except sample class6
0.70 (141) 0.44 (129) 0.47 (138) 0.44 (138) 0.57(142)
0.47
0.64
0.34
0.33
0.61
0.50
0.68
SBRP Deposition ns ns ns ns ns
Depn., land use ns (30) ns (30) 0.21(26) 0.12 0.35(26)
Depn., soil sample 0.52 (30) 0.41 0.38 0.23 (28) 0.92 (30)
class
Depn., all mapped 0.77 0.52 (30) 0.42 (32) 0.35 (32) 0.89 (25)
attributes
Soil chemistry 0.32 - 0.29
Depn., soil chemistry, 0.73 0.44 0.82 0.65 0.85
mapped attributes
except sample class
a For most analyses 143 (NE) or 31 (SBRP) watersheds were used in regressions; the number of watersheds varies among
analyses due to deletion of outliers and/or influence points and is indicated if different from 143 or 31.
b No variables significant in regression equation.
0 Based on sample class area] extent (percent of watershed area).
d Independent variables include deposition, geomorphic and derived hydrologic variables, mapped geology (areal extent),
land use, vegetation, sample class areal extent, and depth to bedrock.
:, except sample class extent dropped and soil chemistry (aggregated to watershed value)
6 Same independent variables as
added
155
-------
NE sample. Variables that characterize reducing environments (e.g., % wetlands, extent of Histosols
and wet soils, presence of beaver ponds) are negatively correlated with suifate concentration and are
positively correlated with percent sulfur retention. Attributes associated with watershed disturbance
(e.g., pits/quarries, urbanization, agricultural activity) are linked to higher suifate concentrations. In the
SBRP, current suifate concentrations in runoff are controlled principally by adsorption in soils (Church
et al., 1989, Section 9,10), with the result that deposition does not directly control surface water
suifate and is not significantly correlated with stream suifate concentrations. In the SBRP, suifate
concentrations and retention are most closely linked to those variables that characterize suifate
adsorption by soils. These include the extent of high- or low-adsorbing sample classes, soil mass,
and soil chemical variables directly or indirectly linked to suifate adsorption.
More generally, we can summarize the major results of the Level I analyses for the DDRP
sample of watersheds in the NE and SBRP as follows:
• Major watershed disturbances, such as quarries and urbanization, are associated not only
with e!evated suifate concentrations, but also with increases in base cation concentrations
and ANC.
• In both the NE and SBRP, agricultural land use and/or the extent of "open dry" vegetation
(i.e., nonforested, undeveloped land other than wetlands, a classification that encompasses
most agricultural land) are associated with high pH, ANC, and base cations and with ele-
vated suifate concentrations. The high pH, ANC, and base cations could reflect either the
preferential use of fertile soils for agricultural activities or anthropogenic amendments, such
as lime, to those soils.
• Surface water chemistry is related to soil depth and particle size in a predictable manner;
suifate retention in SBRP soils and high base cation output are associated with deep, rela-
tively fine-textured soils, whereas high suifate concentration (low retention) and low base
cations and ANC are linked to predominance of shallow and/or coarse-textured soils on
watersheds.
• Regression analyses using only soil chemistry data (aggregated to watershed values)
resulted in values of R2 substantially lower than regressions based on combinations of
deposition and landscape variables and also lower than regressions for soil chemistry data
used in combination with deposition and with mapped watershed attributes. Soil chemistry
variables associated with base cation concentration and pool size (pH, CEC, % base satura-
tion) are positively related to pH, ANC, and base status of surface waters; suifate adsorption
variables for the soil (suifate pools, isotherm coefficients) are positively related to surface
water suifate concentration.
• The soil sampling classes developed for the NE and SBRP, which were delineated to. reflect
variability in soil chemistry, drainage, texture, thickness, and parent material, are effective for
describing landscape units. Based on the observation that regression analyses using just
sample classes and deposition resulted in values of R2 comparable to those developed in
integrated analyses using all landscape variables, we conclude that the sample classes
capture the diversity of both chemical and landscape attributes. The values of R2 for these
analyses often were higher than those resulting from analyses based solely on aggregated
soil chemistry data and (in some cases), were comparable to the even higher values of R2
for analyses using deposition, soil chemistry, and all mapped attributes (except sample
class).
156
-------
Results of a parallel set of regression analyses, conducted using data for riparian buffer
zones, were inconsistent. In some cases, regressions using attributes within the buffer zone
resulted in values of R2 higher than those for the entire watershed, but in other cases values
of R2 for the buffer were comparable to or lower than those for the entire catchment. The
inconsistent results for analyses using riparian buffers may be caused by measurement error
(i.e., the spatial resolution and minimum area of map unit delineations).
Overall, the results were consistent with current hypotheses regarding controls on surface
water chemistry and with the processes and relationships incorporated into watershed
chemistry models. Results do not suggest any major omissions or inconsistencies with the
processes incorporated in watershed models used in our Level II and III analyses.
157
-------
-------
SECTION 9
LEVEL II ANALYSES - SINGLE FACTOR RESPONSE TIME ESTIMATES
9.1 INTRODUCTION
Although a number of watershed processes influence surface water chemistry (e.g., see
summary reviews in Church and Turner, 1986; Sections 2 and 3 of Church et al., 1989), changes
related to a handful of processes probably represent the major controls on long-term changes in
watershed response to sulfur deposition. The MAS Panel on Processes of Lake Acidification (NAS,
1984) focused on sulfate adsorption and base cation exchange by soils as critical time-variant
processes that might contribute to a delayed response to acidic deposition. In its focus on sulfate
adsorption and cation exchange, the NAS panel recognized scientific uncertainties concerning the
present and potential long-term roles of the two processes. Mineral weathering and cation uptake by
vegetation are examples of two rate-limited processes; the magnitudes of which are likely to change
slowly, if at all, over the 50-year period of the DDRP projections. Adsorption and exchange, however,
are capacity-limited processes. As sulfate adsorption sites become occupied or as exchangeable
cations are leached from the soil, the buffering capacities of these two processes decrease monotonic-
ally, increasing the probability of surface water acidification. The time frame of such changes can vary
widely as a function of soil physical and chemical properties. In watersheds with thin or very coarse-
textured soils, buffering of acidic deposition by adsorption or exchange can be very limited and some
systems can respond almost immediately. Alternatively, watersheds with deep soils and high adsorp-
tion capacities and/or large exchangeable base cation pools might experience significant changes in
soil leachate chemistry in response to high acidic deposition loadings only after decades to centuries.
This section presents results of Level II analyses, which involve simulations of the temporal
response of individual watershed processes considered in isolation. We have limited analyses in this
report to an evaluation of the role of sulfate adsorption and cation exchange as mechanisms contrib-
uting to delays in surface water acidification for watersheds in the M-APP Region and, for comparison,
a summary of results for the NE and SBRP previously reported in Church et al. (1989). Analyses are
based on models that consider only adsorption or exchange, and that consider only changes within
the upper regolith {< 2 m for the M-APP and < 1.5 and 2.0 m in the NE and SBRP, respectively).
These analyses assess the influence of adsorption and exchange on soil and/or surface water chemis-
try at the present time, and project probable future changes in the contributions of adsorption and
exchange. The analyses use soil chemistry data collected during the DDRP soil surveys and model-
based projections of (1) future changes in sulfate mobility controlled by sulfate adsorption, and (2) two
independent projections of exchange-mediated changes in base cation leaching and changes in soil
pH and/or cation exchange pools. By considering cation exchange but not mineral weathering (i.e.,
base cation resupply), the exchange models provide a worst-case projection of future changes in soil
and solution chemistry.
The analyses presented in this section evaluate the temporal response of single processes to
acidic deposition. Because they assess those processes in only a portion of the watershed (i.e., the
upper 1.5 - 2.0 m of watershed soils), the model projections should not be interpreted as integrated
158
-------
projections of watershed response time. Rather, these analyses represent a set of bounding estimates
of the relative importance, now and in the future, of the potential role of adsorption and exchange as
delay mechanisms. Analyses presented by Church et al. (1989, Section 7) showed that processes
other than adsorption are of limited importance in mediating sulfate mobility in most NE and SBRP
watersheds, especially in upland catchments such as those considered in this report for the M-APP
Region. Similarly, although several processes other than exchange and weathering affect cation
cycling in terrestrial ecosystems, exchange is generally regarded as the rate- or capacity-limited
process the dynamics of which are most likely to be significantly affected by acidic deposition. Level
111 modelling (Section 10) provides projections of changes in future surface water chemistry based on
integration of adsorption and exchange with other processes.
9.2 EFFECTS OF SULFATE ADSORPTION ON WATERSHED SULFUR RESPONSE TIME
9.2.1 Introduction
As noted in Section 9.1 and as discussed in previous DDRP results for the NE and SBRP
(Church et al., 1989), a major DDRP focus is on sulfate, which is the principal anion in acidic deposi-
tion and the major mobile anion contributing to surface water acidification in the eastern United States.
Because the extent and duration of sulfate retention within watersheds varies widely within and among
regions, depending on deposition history and on soil physical and chemical properties, variability in
sulfate retention by soils represents one of the most important factors influencing the rate of watershed
chemical response (i.e., changes in ANC) to acidic deposition (Johnson and Cole, 1980; Galloway et
al., 1983; NAS, 1984).
At the start of the DDRP, there was believed to be a strong north-south gradient in sulfate reten-
tion characteristics of soils. Soils in the glaciated northern areas of North America were perceived to
have low sulfate adsorption capacity, resulting in negligible sulfate retention by soils, and watershed
sulfur budgets at or near steady state. In contrast, watersheds in the southeastern United States were
characterized by high net sulfur retention, which was attributed to the moderate to high sulfate adsorp-
tion capacities of deep, highly weathered soils in the region (NAS, 1984). Little was known, however,
about either adsorption characteristics of soils or watershed sulfur budgets for the M-APP Region. Soil
analyses conducted as part of the DDRP in the NE and SBRP demonstrated strong regional differ-
ences in sulfate adsorption characteristics that are consistent with this paradigm. Similarly, site-
specific and regional analyses of watershed sulfur budgets (Rochelle et al., 1987; Rochelle and
Church, 1987; Church et al., 1989) have confirmed north-south differences in regional sulfur budgets,
and in addition have shown a range of retention in M-APP watersheds (Section 7).
Along with budget data showing a range of sulfate retention by M-APP watersheds, there are
scattered data describing sulfate dynamics for watersheds within the region. Monitoring data for
streams at Fernow, West Virginia (D. Helvey, pens, comm.), and at White Oak Run and Deep Run,
Virginia (Ryan et al., 1989), document increases in streamwater sulfate concentrations of about 1-2
/teq/L/yr. Modelling of the White Oak Run catchment projects that sulfate concentration will increase
from levels in the mid-1980s of around 80 ,«eq/L to over 200 fieq/L during the next 50-60 years (Cosby
et al., 1985b,c 1986b). The regional representativeness of these few sites is unknown; the modelling
159
-------
activities described in this section provide the first regional assessment of surface water sulfate
dynamics for the M-APP Region.
9.2.2 Section Objectives
This section assesses changes in sulfate mobility in DDRP watersheds and regions that are
attributable to sulfate adsorption (and desorption) by soils. The importance of processes other than
adsorption on sulfate mobility has been previously assessed for watersheds in the NE and SBRP, and
these other processes were found to be of little importance in most DDRP watersheds. The overall
goal of the Level II analysis is to assess the importance of sulfate adsorption as a process contributing
to delays in surface water acidification within study regions. Specific objectives of Level II sulfate
analyses for the M-APP Region, presented here, are identical to those for the previous NE and SBRP
analyses:
• Characterize and compare sulfate pools and sulfate adsorption capacities of soils within and
among study regions.
• Estimate the response time of soils in DDRP watersheds to changes in sulfur deposition
using an adsorption-based model.
• For watersheds not presently at steady state, forecast time to steady state under current
deposition loadings. In addition, for all watersheds, forecast response time to future
changes in deposition. Results for individual watersheds will be extrapolated to regional
forecasts.
• Summarize the contributions of sulfate adsorption to delays in surface water acidification
resulting from historic or future changes in deposition.
The results of these analyses will also provide data for evaluating and comparing the relative
importance of sorption and other processes considered by the DDRP models (e.g., cation exchange),
although comparisons are not made within this section.
It is important to recognize that procedures and models used for this analysis consider only
sorption processes. Processes affecting watershed chemistry, other than those involving sulfate, are
not considered, and watershed conditions and fluxes (except sulfur deposition) have been held con-
stant for the duration of model projections. It is equally important to recognize that the forecasts and
estimates of time to steady state made here apply only to sulfate. Although change in sulfate mobility
is one of the principal factors driving change in surface water cations and ANC, processes that don't
involve sulfur also play crucial roles in the dynamics of those ions. Rates of change in ANC, and
especially potential time to zero ANC, are not necessarily coincident with time to steady state for
sulfur. Systems can reach zero ANC before, concurrently with, or after time to sulfur steady state.
The relationship between changes in sulfate and ANC are characterized and discussed as part of the
Level 111 analyses in Section 10.
Although the model projections included in this discussion extend beyond 50 years, attention
should be focused on response over the first 50 years of simulations. Projected changes beyond the
160
-------
50-year period have been included for several reasons. First, in many systems, sulfate response is
projected to occur over time frames exceeding 50 years, in some cases for a period likely to last for
well over a century. Second, including long-term projections indicates the probable magnitude of
changes occurring after the 50-year time frame, until steady state is reached. Also, for the scenarios
of altered deposition, simulated changes in deposition continue for much or all of the 50-year
response period. It is difficult, if not impossible, to effectively assess the response to those changes
by considering only a 50-year, rather than a long-term, system response.
9.2.3 Approach
Level II sulfate analyses are based on model-based projections of future sulfate dynamics in 36
DDRP watersheds within the M-APP Region, with results extrapolated to a regional target population.
Procedures and data sources are comparable to those used for similar DDRP sulfate analyses for the
NE and SBRP. Soil mapping and chemistry data generated by the DDRP Soil Surveys (Section 5.4,
5.5) were used to make the projections. The principal soil variable for these analyses was sulfate
adsorption isotherms generated for individual soils collected in the surveys, then aggregated to the
watershed level. Forecasts were made using a modified version of the sulfate subroutine in the Model
of Acidification of Groundwater in Catchments (MAGIC) (Cosby et al., 1985a, 1986b). A detailed
description of the model, data sources, underlying assumptions, etc., was presented in Church et al.
(1989); key points are repeated below.
9.2.3.1 Model Description
AH model projections described in this section were made using a modified version of the sulfate
subroutine from the MAGIC model, which was originally developed for stream systems in the Virginia
Blue Ridge (Cosby et al., 1985a,b; 1986b). The model uses a deterministic, mass-balance approach
that considers only adsorption as a sulfur retention process by soils (Cosby et al., 1985b,c; 1986b).
Sulfate partitioning between dissolved and sorbed phases is defined by a hyperbolic (Langmuir) iso-
therm. The original MAGIC subroutine has been modified to accommodate multiple soil horizons (up
to 10, although 3 were used for this study). Soil horizons are treated as a series of continuously
stirred tank reactors (CSTRs); all precipitation and sulfur deposition (wet and dry) inputs are made to
the top mineral soil horizon (organic horizons are not considered in the model, because sorption is
negligible in the O horizon). Evapotranspiration implicitly occurs in the top soil horizon. All flow is
then routed sequentially through each soil horizon. Data are input to the model using annual time
steps. The projected surface water sulfate concentration is defined by (set equal to) the equilibrated
solution sulfate concentration in the lowest soil horizon. Because sorption is essentially an instan-
taneous process, reaction kinetics are not considered and equilibrium between solution and sorbed
phases is assumed to occur in all cases.
For these analyses, model simulations were run starting 140 years prior to the base year of 1986
[i.e., the year of the EPA National Stream Survey - Phase I (NSS-I), which included Mid-Appalachian
streams (Kaufmann et al., 1988)]. Comparison data for NE lakes (Linthurst et al., 1986) and SBRP
streams (Messer et al., 1986) employ base years of 1984 and 1985, respectively. Soil and stream-
water sulfate concentrations were set at the start of simulations by assuming both to be at steady state
161
-------
with respect to deposition (low historic levels). Simulations were run for a sufficient time into the future
to allow systems to reach steady state. For most M-APP watersheds, simulations were run for 140
years; in a few cases with long response times, simulations were extended to 300 years.
in calibrating the model for M-APP systems, as described in Section 9.2.3.5, simulated response
was constrained to compare favorably with two kinds of surface water data. First, the distribution of
simulated sulfate concentrations (and of percent sulfur retention) for the base year had to compare
closely to the observed distribution of streamwater sulfate. Secondly, because many trajectories of
simulated response can match current sulfate concentration, we also required simulated rates of
change for sulfate to be comparable to observed rates for streams monitored within the region. We
present and discuss these comparisons in Section 9.2.4.3.
As in previous analyses for the NE and SBRP (Church et al., 1989), we define response times as
times for projected sulfate concentrations to reach 95 percent (or for systems with decreasing sulfate
concentration, 105 percent) of steady state. We use this approach for two reasons. First, model
projections approach steady-state concentrations asymptotically, such that true steady state is not
reached. Second, and more important from a practical perspective, given the measurement uncertain-
ties and interannual variability in precipitation, runoff, deposition, and surface water sulfate
concentration, the difference between 95 and 100 percent of steady-state concentration is not
distinguishable.
9.2.3.2 Data Sources
Input requirements for the sulfur model include current sulfur inputs and outputs (precipitation,
runoff, total sulfur deposition, and sulfate concentration in runoff), scenarios of historic and future
sulfur deposition, and soil variables to describe sulfate partitioning and adsorption capacity (adsorption
isotherms, soil mass). Data requirements and sources are comparable-to those for similar DDRP
. sulfate modelling in the NE and SBRP.
The precipitation and deposition datasets used here are analogous to the typical year datasets
generated for DDRP analyses in the NE and SBRP; generation of these datasets is described in Sec-
tion 5.6.3. Runoff estimates, based on interpolation of 30-year average USGS runoff maps, were
generated as described in Section 5.7.1. Data for current streamwater sulfate concentrations are
taken from the results of EPA's NSS-I in the Mid-Atlantic and southeastern United States (Kaufmann et
al., 1988), conducted as part of the EPA's National Surface Water Survey.
Historic sulfate deposition in the region is unknown; initial sulfate inputs (year -140) were
estimated as 5 percent of current deposition. We based estimated sulfur deposition between initial
and base years (1846-1986) on SO2 emission estimates of Gschwandtner et al. (1985) for Federal
Region III (Delaware, Maryland, Pennsylvania, Virginia, West Virginia); we used linear interpolation
between the initial simulation year and 1900. The historic emission pattern served as a scaling factor
for each watershed, a procedure that implicitly assumes a constant relationship between regional
emissions and site-specific deposition over the last 140 years.
162
-------
Three scenarios of future sulfur deposition were used for the M-APP Region. The first scenario
(A) calls for constant deposition through the 140-year simulation period (Figure 9-1). The second and
third scenarios are characterized by reduced deposition in the future. Both start with constant deposi-
tion at current rates for years 1 to 5, followed by a linear reduction in deposition, to 50 percent of
current loadings, between years 6 and 15. For the second scenario (B), future deposition is main-
tained at 50 percent of current loadings from year 15 through the end of the simulation period. The
third scenario (C) allows a linear increase in deposition, from 50 to 80 percent of current loadings
between years 15 and 50, followed by constant deposition at 80 percent of current input from year 50
to the end of the simulation period. All comparisons of regional projections among the three DDRP
regions are based on scenarios of constant future deposition.
Mapping of soils and quantification of the areal extent of various soils on DDRP watersheds are
described in Section 5.4. Sampling and chemical/physical analyses of soils are described in
Section 5.5. For each mineral soil horizon, sulfate adsorption data were used to compute adsorption
isotherms, which were then aggregated, using mass weighting (computed from horizon thickness, bulk
density, and coarse fragment content) to obtain sample class and watershed values. Procedures for
derivation of adsorption isotherms and for aggregation of adsorption data are identical to those used
for previous DDRP analyses in the NE and SBRP, which are described by Shaffer and Stevens (1991)
and Section 9.2. of Church et al. (1989).
During the process of sample class development and sampling site selection for the Mid-
Appalachian Region, we recognized that in those sample classes that occurred across the entire
region, the potential existed for significant within-class differences in soil chemistry related to
geography (e.g., glaciated parts of the northeastern Pennsylvania Plateau, and Valley and Ridge
physiographic provinces) and/or to gradients in current and historic rates of acidic deposition. If such
differences exist, but were obscured by a forced combination of data during aggregation, the quality of
subsequent model simulations might be compromised. As part of M-APP analyses, we have analyzed
spatial trends for soil sulfur data. Results indicate that, when data for all M-APP pedons are consid-
ered, there are significant differences in the spatial distribution of relevant sulfur variables (e.g.,
extractable sulfate, adsorption isotherm slope), but they result principally from differences in the spatial
distribution of sampling classes within the region. With few exceptions, within-class spatial patterns
are not significant The single exception, sample class END, has low values for adsorption isotherm
slope and SO4 -PO4, in the Valley and Ridge province, but covers no more than 20 percent of any
watershed in that subregion. If data were subdivided and reaggregated to watershed scale, for
sample classes that are geographically widespread, there would be only very minor differences in the
reaggregated data. Available results indicate that the current sample class structure and aggregation
are reasonable arid are well-suited to our regional analyses.
9.2.3.3 Model Assumptions and Limitations
Several critical assumptions are encompassed by the choice of model and the methods of data
collection. These in turn impose limitations on the scope and interpretation of model projections. Key
model assumptions and their implications for data interpretation include:
163
-------
FUTURE SULFUR DEPOSITION SCENARIOS
DDRP MID-APPALACHIAN REGION
X
D
LL
nr
ID
LL
_l
=)
CO
LU
h-
_J
LU
DC
120-
oo -
80 -
.
RO-
40-
20-
-
\
\
\
* s — — — — —
\
,'
\
*
\ s
Scenario A
Scenario B
Scenario C
I 1 i • i • i • i • i
0
25 50 75 100 125
YEAR OF SIMULATION
150
Figure 9-1. Deposition scenarios used for modelling future sulfate dynamics in soils of
DDRP M-APP watersheds.
164
-------
• Sorption is the only watershed "process affecting sulfate mobility and watershed response -
time. As noted previously, this limitation provides the most effective means of assessing the
significance of adsorption by soils as a process delaying surface water acidification. To the
extent that other terrestrial processes sequester or generate sulfate on a net basis, model
projections will under- or overestimate the.time and/or magnitude of the projected response.
As noted earlier, the net role of other processes in most DDRP watersheds is believed to be
small, especially in upland systems that lack extensive wetlands or flooded soils.
• The analytical approach used to define sulfate partitioning in the soil (hyperbolic isotherms
determined by batch equilibrium methods using air-dried soils) adequately describes sulfate
partitioning by soils under field conditions. A methods comparison study by Hayden (1987)
supports the use of hyperbolic isotherms and batch equilibrium methods. Recent analysis of
soil handling effects (Comfort et al., 1991) indicates that the effects of air drying on
adsorption increase in proportion with organic content. Soils with low organic content were
not significantly affected by drying. Because most adsorption occurs in soils with low
organic content (organics block anion adsorption), the practical importance of drying effects
appears to be negligible.
• Soil and watershed conditions influencing adsorption (e.g., soil pH, Fe, Al, and organic
content) are assumed to be constant for the duration of model projections. pH is probably
the most important variable of concern, since adsorption is strongly pH dependent. Level II
cation exchange models project moderate decreases in soil pH during the simulation period,
which would be expected to lead to modest increases in sulfate adsorption capacity. The
quantitative relationship between soil pH and adsorption is not well defined (and thus is not
incorporated into the model), but large effects on model projections would not be expected.
• Hydrologic routing is simple; flow through the soil is represented by a series of continuously
stirred tank reactors (CSTRs), and all flow is routed sequentially through each horizon. The
* perfect hydrologic contact represented by a simplified flowpath used here does not
realistically reflect lateral flow, macropore flow, etc., that occur in the soil. Data needed to
objectively define routing coefficients for alternate flowpaths are lacking, however, so
alternate routing was not incorporated into model projections. The effect of modelling flow
to bypass upper or lower soil horizons under natural conditions would be projections of
higher initial sulfate leaching (part of the input signal would not be attenuated by sorption on
the soit), but a more gradual (in terms of change in concentration with time) subsequent
sulfate response. The responses projected here represent an upper bound on initial
response time, assuming complete contact between the soil and flow through the soil, and a
lower bound on time to steady state.
• Because the model runs on an annual time step and uses identical precipitation, runoff, and
sulfur deposition data from year to year, projections do not capture the effects of interannual
variability of natural systems. The lack of such variability in the DDRP projections is
recognized, but should have little effect on the primary objective of these analyses, i.e. to
project long-term changes attributable to chronic sulfur deposition. If there were any long-
term trends in precipitation or runoff, they would not, of course, be represented by model
projections.
9.2.3.4 Adsorption Data
Data describing sulfate partitioning by soils, which are used to develop partitioning functions
(isotherms) of sulfate adsorption capacity of soils, were generated for individual soils as part of the
DDRP soil survey (Section 5.5). Adsorption isotherms were developed for each soil, then aggregated
165
-------
to sample class and to watershed values using mass weighting procedures. Johnson et al (1990)
have described the conceptual objectives of aggregation for DDRP soil chemistry data; actual
aggregation procedures are described by Shaffer and Stevens (1991) and summarized following the
discussion of adsorption isotherms immediately below.
Adsorption isotherms were generated from data for soil-water slurries equilibrated with six
different amounts of sulfate (0, 2, 4, 8, 16, and 32 mg S/L) designated as SO4_0, SO4_2, etc., in
Section 5.5.4.2.1. For each of the six samples, net sulfate adsorbed by the soil was computed from
the change in dissolved sulfate. For example, for the 8 mg/L sample:
SO4_8n = (SO4_8, - SO4_8f)
L
S
(9-1)
where: SO4_8j
SO4_8f
L
S
= dissolved sulfate concentration before equilibration (^eq/L)
= final dissolved sulfate concentration after equilibration (aeq/L)
= volume of liquid (~ 0.050 L)
= mass of volume of soil (~ 0.010 kg)
Raw data points were then fit to an extended Langmuir isotherm ("extended" by addition of a third
variable to describe the non-zero Y-intercept) (Hayden, 1987), the partitioning equation taking the
form:
Ec =
B1 * C
B + C
+ B,
(9-2)
where: B1 = maximum sulfate adsorption (meq/kg)
B2 = half saturation constant (aeq/L)
B3 = Y-intercept (meq/kg)
C = dissolved sulfate (weq/L)
Ec = net adsorbed sulfate at [SO-/'] = C (meq/kg)
The three parameters (B^ B2, and B3) were estimated by a nonlinear least squares procedure
(Fletcher and Powell, 1963; Hayden, 1987) to minimize the sum of squares function. Examples of raw
data and resulting fitted isotherms are shown by Church et al. (1989, their Figure 9-1), and Shaffer and
Stevens (1991, their Figure 1).
Several approaches were evaluated for aggregating isotherm data from individual soils; the
"aggregate isotherm" approach of Shaffer and Stevens (1991) was selected because it has been
shown to generate effective weighted averages for these nonlinear functions. This procedure involves
three steps as follows: (1) after fitting isotherms for individual soils, compute concentrations of net
adsorbed sulfate corresponding to several concentrations of dissolved sulfate (0, 10, 25, 40, 75, 125,
200, 500, 1000, 2000 ^eq/L) for each soil, (2) for each value of dissolved sulfate, compute a mass
weighted average of the corresponding concentrations of net adsorbed sulfate for the soils in each
aggregation group, then (3) fit a new isotherm to the set of weighted averages. The resulting isotherm
166
-------
was defined as 4he aggregate isotherm and was used to describe sulfate partitioning for that group of
soils. For routine applications, isotherm aggregation data were aggregated to the master soil horizon
(A/E, B, or C), using a three step procedure: (1) individual soil (sub)horizons to master horizon within a
pedon (mass weighting), (2) pedon (by horizon) to sample class (by horizon), using mass weighting,
and (3) sample class to watershed (by horizon), using mass and area weighting of each sample class
occurring on each watershed. Missing data were assigned the aggregate value for other subhorizons
In their respective pedon; for soils having only one subhorizon within a master horizon, the sample
class average for that horizon was assigned to missing values. For routine uses, data for soil master
horizons were used directly in the model and were not further aggregated.
Because the aggregation approach was not conducive to direct computation of parameter
uncertainty (i.e., computing a mean and standard deviation for coefficients), we used an alternate
approach. Uncertainties for the original isotherm fits were retained, then a Monte Carlo procedure was
used during each step of aggregation to generate estimates of uncertainly in aggregated coefficients
at the sample class and watershed level. Uncertainty in sulfate isotherms was propagated through the
aggregation procedure using the Monte Carlo technique described in Section 6.3, with steps similar to
those used for aggregation of other variables. We repeated the aggregation from individual sub-
horizons to sample class master horizon 100 times, each time selecting a randomly perturbed set of
coefficients for each subhorizon isotherm. The perturbation of B1 was selected first from a normal
distribution with a standard deviation obtained from the residual sum of squares and the inverse
Hessian matrix from the initial nonlinear least squares fit. The perturbed value of B^ along with the
correlation of B1 and B2 from the inverse Hessian matrix, was then used to estimate the conditional
expectation of B? given Br This conditional value was then perturbed by a value drawn from a normal
distribution with the conditional standard deviation of B2 given Br A similar procedure was used to
perturb B3, except that the mean and variance were adjusted for both B1 and B2. The mean values,
standard deviations, and correlation matrix of the coefficients were summarized at the sample class
level. These values were then passed to the watershed level aggregation algorithm. The uncertainty
calculation for watershed aggregation was conducted using the same procedure as that described
immediately above (this paragraph) for sample class uncertainty estimates, except that the correlations
were derived from the sample class Monte Carlo study rather than from an inverse Hessian matrix.
9.2.3.5 Calibration and Evaluation of Aggregated Data
Three approaches were used to evaluate aggregated soil sulfate data and to calibrate data for
model simulations; all used isotherm data aggregated to watershed level. First, equilibrium soil
solution sulfate concentrations (the dissolved sulfate concentration at which net adsorbed sulfate
equals zero; see Equation 9-2) for B horizon isotherms, aggregated to watershed values, were com-
pared to streamwater sulfate concentrations. If adsorption by the soil were the sole process influencing
sulfate mobility, if lab isotherms precisely represented partitioning under field conditions, and if
aggregation procedures were perfect, a 1:1 correlation between soil and lake/stream sulfate concentra-
tions would be obtained. Realistically, due to contributions of factors such as hydrologic routing,
heterogeneity of natural soils, uncertainties introduced by soil sampling and analysis, and effects of
regionally focused (rather than watershed by watershed) sampling and aggregation, a high correlation
between soil solution and surface water sulfate was not expected. The purpose of this comparison
167
-------
was to evaluate whether the two sets of values had comparable ranges and whether major biases exist
that might invalidate the entire approach.
The second and third approaches to evaluation of isotherm data use model simulation results,
and compare model-based data to measured stream chemistry. Isotherm and soil data were cali-
brated by comparing projected sulfur retention in the base year to the percent retention computed as
described in Section 7. Simulations were run starting 140 years before the base year, and model
simulations were not initialized in any way that constrained year 0 projections to match measured
chemistry in the base year. The comparability (or lack thereof) between measured and observed
sulfate retention for the base year thus provides a check point for simulations; isotherm data were
calibrated by scaling coefficients (B2, or B2 and B1 together if necessary) such that the mean (for all
watersheds) percent sulfate retention for measured and simulated sulfate were approximately equal.
Calibration was accomplished by scaling the coefficients for all watersheds by the same amount,
resulting in simulations that are unbiased for the population but not matched on a watershed-by-
watershed basis.
The final check on isotherm data was to compare measured and modelled rates of change in
streamwater sulfate. Rates determined from monitoring studies were compiled and compared to rates
generated from model simulations. Because field data were available for only a few field sites, this
comparison provides only a rough qualitative check, but it does insure that the trajectory of
simulations (i.e., the rate of increase for sulfate with time) is not unrealistic.
9.2.3.6 Target Populations for Model Projections
Because the modelling watersheds used for DDRP were selected within the probabilistic
framework of EPA's National Surface Water Survey, results from the sample watersheds can be
extrapolated to an explicitly defined regional target population (Linthurst et al., 1986; Kaufmann et al.,
1988). Except as noted, all results presented for the M-APP Region, and all comparison results from
previous NE and SBRP analyses, represent lake or stream target populations within their respective
regions. In the M-APP Region, 36 watersheds were modelled within the DDRP, representing a popula-
tion of 5,496 stream reaches within the region (Kaufmann et al., 1988). This population represents a
subsample of the National Stream Survey watersheds having (1) ANC of <200 fieq/L, (2) sulfur
budgets for which sulfur output was not significantly higher than input (i.e., no major internal
watershed sulfur sources), (3) catchment areas of 30 km2 or less, and (4) no evidence of acid mine
drainage (as evidenced by streamwater chemistry, topographic maps, or visible disturbance, or any
combination of these). Sample and target populations for NE lakes and for SBRP streams were
described in detail in Section 9.2 of Church et al. (1989). For the NE, simulations were made for 131
lakes representing a target population of 3,314 lakes. Thirty-four stream systems were modelled in the
SBRP, representing a population of 1,492 stream reaches.
168
-------
9.2.4 Results
9.2.4.1. Summary of Soil Adsorption Data
Sulfate adsorption data for soils in the M-APP are summarized in Table 9-1; for purposes of
comparison, data for the NE and SBRP are also included. For most of the listed parameters, soil
concentrations or pools for M-APP soils are intermediate between those of the other two regions.
Mass-weighted average concentrations of phosphate-extractable sulfate for M-APP soils range between
25 and 57 mg S/kg; for each horizon, the concentrations are higher than for the equivalent NE soils
but lower than for corresponding SBRP soils. Adsorption isotherm coefficients [adsorption maxima
(B1) and half-saturation constant (B2)] for M-APP soils are also intermediate in value, although the
interdependence of these coefficients makes the relative differences difficult to interpret. Figure
5-10 indicates the variability of these data on a sample class and pedon basis.
If we compare weighted average isotherm slopes (partitioning coefficients between sorbed and
solution phase) instead of individual coefficients, and consider the B and C soil horizons in which most
adsorption occurs, the average slopes for M-APP soils are two to four times higher (slopes of 6.8 and
7.35 versus 3.15 and 1.31) than for NE soils, but only about one half the values for soils in the SBRP.
Similar ratios are observed for the amount of sulfate that can be adsorbed per unit mass of soil at
steady state. The differences among the three regions are even more pronounced if we consider the
pools of adsorbed sulfate (at sulfur steady state) per unit watershed area. Whereas the average pool
in the M-APP is almost 5 times that of NE soils (17.1 kg/ha vs 3.7 kg/ha) the pool for SBRP soils (55.4
keq/ha) is over 3 times as large as the M-APP pool and about 15 times that of NE soils. This relation-
ship occurs because average soil mass is greater in the SBRP than in the M-APP, and the B horizon
mass is greater in the M-APP than in the NE.
The differences among sulfate retention characteristics for soils in the three regions reflect differ-
ences in soil age, parent material, and other soil chemical properties. NE soils are much younger
than those in the other regions, (around 14,000 years, versus millions of years) and thus are thinner.
They are characterized by lower amounts of inorganic hydrous oxides and clays than soils in the other
regions. Furthermore, the NE soils have higher concentrations of organic carbon, which interferes with
or "blocks" sulfate adsorption, than soils in the M-APP or the SBRP. In comparison to SBRP soils,
adsorption in the M-APP is lower for several reasons: (1) soils in the M-APP are shallower; thus (all
other things being equal) they have less soil mass and surface area than SBRP soils, (2) most soils in
the M-APP have developed from sedimentary bedrock, from which the readily weatherable primary
minerals were lost during a previous weathering episode, and (3) the more slowly weathering sedi-
mentary rocks in the M-APP experience less dissolution of Fe and Al from aluminosilicate minerals,
providing less substrate for formation of secondary clay mineral phases and hydrous oxide coatings.
9.2.4.2. Evaluation of Aggregated Data
As described in Section 9.2.3, several comparisons between measured field variables and either
adsorption isotherms or model simulation data were used as checks on the reliability of model inputs
and outputs. The first such check compared streamwater sulfate concentrations to equilibrium soil
169
-------
Table 9-1. Summary Comparison of Adsorbed Sulfate and Adsorption Capacities of Soils on
Watersheds in the DDRP Target Populations in the M-APP, NE, and SBRPa
Soil
Region Horizon
M-APP A
B
C
NE A
B
C
SBRP A
B
C
Soil
Mass
g/cm2
10.7
78.0
26.1
10.5
55.1
68.9
17.1
90.5
68.2
isotherm Coefficients13 Isotherm
SO4-PO4 B1 B2 slope©
mg/kg meq/kg /*eq/l_ lOO^eq/L
25.0
57.2
33.1
22.0
35.1
13.5
47.7
88.9
42.0
3.23 1788 1.71
5.25 577 6.80
4.82 457 7.35
3.12 1361 2.03
3.57 945 3.15
1.60 1058 1.31
4.38 1093 3.50
7.55 183 17.10
6.25 109 14.70
Adsorption at Sulfur
Steady State
meq/kg keq/ha
0.40
1.58
1.68
pedon 2 =
0.26
0.41
0.16
pedon 2 =
0.45
3.23
3.73
pedon 2 =
0.4
12.3
4.4
17.1
0.3
2.3
1.1
3.7
0.8
29.5
25.4
55.4
a Median Values are Listed.
b Coefficients used in equation 9-2.
170
-------
solution sulfate (or ESSS, which represents the concentration of dissolved sulfate at which net
adsorbed sulfate equals zero) for B horizon soils, using soil data aggregated to watershed scale.
Results of this analysis, which assesses whether measured isotherms predict sulfate concentrations
comparable to observed streamwater sutfate concentrations, are shown in Figure 9-2. A scatter plot of
data shows that in comparison to the wide variability in measured streamwater chemistry (55-313
fteq/L), values of ESSS fall within a narrow range, with all but two B horizon soils having values of
ESSS between 131 and 210/teq/L Mean and median sulfate concentrations from the two data sets
are similar, however. Mean and median concentrations for stream water are 152 and 154/teq/L,
respectively, with corresponding values of ESSS for B horizon soils of 164 and 170 fieq/L. The rela-
tively narrow range of values for ESSS is in all likelihood a result of using soil sampling and data
aggregation procedures optimized for regional, rather than site-by-site, soil characterization and model
projections. The regional approach, as discussed by Church et al. (1989), effectively characterizes
central tendencies and broad regional variability of watershed characteristics and surface waters, but
was not intended to (and does not) fully reflect the range of intra-regional variability in soil and surface
water chemistry. The most important result of this comparison is that although the range of values in
the two data sets is different, mean and median values are similar, suggesting that there are not fun-
damental problems or major systematic biases with the isotherm data or with aggregation procedures.
Recall also that the relative heterogeneity of watershed-scale values of ESSS cannot be attributed to
aggregation of geographically diverse data for soils within sampling classes. Analysis of variance for
data within sample class do not indicate systematic differences in soil sulfate between geographic/
physiographic subregions or across deposition gradients.
A second data check, comparing measured stream sulfate to modelled stream sulfate for the
base year of 1986, indicates similar ranges and distributions of sulfate concentrations in the two data
sets. Mean measured and modelled sulfate concentrations are 152 and Isy^eq/L respectively; corre-
sponding mean values of percent sulfur retention are 35 and 34 percent. Site-by-site comparison of
the data (Figure 9-3, A and B) indicates a great deal of scatter. As already noted, however, this result
is not especially surprising given that sampling and data aggregation were designed to characterize
regional, rather than site-specific, characteristics of soils and watersheds. When data are presented to
compare populations, using cumulative distributions (Figure 9-3 C and D), the distributions of meas-
ured and modelled sulfate, for both sulfate concentration and percent sulfur retention, are very similar.
In contrast to previous modelling results for the NE and SBRP, for which data were not adjusted
during the calibration procedure, isotherm data for all M-APP watersheds required significant calibra-
tion to match distributions of observed and simulated sulfate for the base year. Two of the isotherm
coefficients were adjusted, with all values of B1 raised by a factor of three, and all values of B2
adjusted by a factor of 0.75. Results described here show that use of the calibrated isotherm data in
model simulations results in projections that match not only the mean watershed sulfate condition for
the region, but also very closely match the regional distributions of sulfate concentration and percent
watershed sulfur retention.
The final check on the isotherm data used in model projeptions was to compare measured rates
of change for sulfate in stream waters of the M-APP to rates projected by model simulations. Because
there is very limited surface water monitoring data from which long-term sulfate trends can be deter-
mined, and because the representativeness of those trend data is unknown, we cannot make a
171
-------
STREAMWATER VS SOIL SOLUTION SULFATE
DDRP MID-APPALACHIAN WATERSHEDS
350
I 3001
LLJ
ID
CO
z
o
H-
_j
O
CO
250-
200-
o 100-
o
N
DC
O
X
CD
150-
50 100 150 200 250 300
STREAMWATER SULFATE (ufiq/L)
350
Figure 9-2. Equilibrium soil solution sulfate (B horizon data) and measured stream sulfate
concentration for the 36 DDRP M-APP watersheds.
172
-------
MID-APPALACHIAN SULFATE
measured vs modelled - SO4 (u,eq/L)
100 150 200 250 300
MEASURED SULFATE (neq/L)
350
MID-APPALACHIAN SULFATE
36 sites; stream vs model year 0 values
u
KO
0,8
0.6
0.4
0.0
50 100 150 200 250 300 350 400
SULFATE (neq/L)
MID-APPALACHIAN SULFATE
measured vs modelled - % Sulfur Retention
-40
B
.» «
-20
0 20 40 60
CALCULATED % RTN
80
MID-APPALACHIAN SULFATE
36 sites; stream vs model year 0 values
1.0
o
C u
Q_
UJ
> 0.4-
0.2-
O
-20 0 20 40 60 80
PERCENT SULFUR RETENTION
100
Figure 9-3. Measured and modelled sulfate concentrations and percent sulfur retention for
the 36 DDRP M-APP watersheds.
173
-------
rigorous comparison. Model projections of sulfate change are, on average, somewhat higher than
reported rates of change for surface waters in the region (Table 9-2). Reported values fall well within
the range of model projections, however, indicating that the model values are not unreasonable esti-
mates of change for sulfate in surface water sulfate.
9.2.4.3. Base Year - Measured Sulfate and Model Projections
As suggested by results presented in the preceding section, measured sulfate concentrations
and model projections of the response of M-APP stream systems to atmospheric sulfate deposition are
indicative of a diverse region with a wide range of surface water sulfate concentrations and response
times. Cumulative distributions of sulfate concentration and of percent sulfur retention for the target
population (5496 stream reaches), based on measured and modelled base year sulfate for the 36
sample stream reaches in the region, are shown in Figure 9-4 and are summarized in Table 9-3.
Because data presented in this and subsequent plots have been weighted to reflect the unequal
sampling probabilities of streams in the National Stream Survey (Kaufmann et al., 1988), mean and
median values of sulfate for the population are different from those for only the sample watersheds,
and the plots for the population differ somewhat in appearance from those for the 36 watersheds
shown in Figure 9-3.
For the target stream population, measured sulfate concentrations range from 55 to 312/ieq/L,
with a mean of 148 and median of 153,ueq/L The calculated percent sulfur retention in M-APP
watersheds (computed using NSWS stream sulfate with precipitation, runoff, and sulfur deposition
estimates described in Section 5.6) is also highly variable, with a mean of 37 percent retention and a
range of -21 to 74 percent. Base year sulfate concentrations projected by the model range from 40 to
365 /
-------
Table 9-2. Rates of Increase for Sulfate in Stream Systems In the M-APP Region, Comparing
Model Projections to Measured Rates of Increase
Site
36 DDRP M-APP
streams
Deep Run, VA
White Oak Run, VA
Fernow, WV, ws #4
Watershed #4
Period of
Record
model-based
estimates
1980- 1987
1980- 1987
1970 - 1985
Stream SO4a
(weq/L)
55 - 313
107
84
85-90
Rate of SO4
Increase
(weq/L/yr)
median = 2.20
range -0.1 - 4.6
Q1 = 1.70
Q3 = 3.53
2.2
2.0
1.0
Reference
this study
Ryan et al., 1989
Ryan etal., 1989
D. Helvey, U.S.
Forest Service
(pers. comm.)
Madison Run, VA
1968, 1982
70
1.3
USGS, 1969,
1970; Lynch and
Dise, 1985
Concentration In last year of record
175
-------
SULFATE CONCENTRATION
1.0
0 50 100 150 200 250 300 350 400
SULFATE (neq/L)
1.0'
z;
o
Hf\ Q ,
«*- w«o
D_
O
O 0.6
Q-
LU
> 0.4
ID
O
0.2-
0.0
PERCENT SULFUR RETENTION
Calculated
Modelled
-40 -20 0 20 40 60 80
SULFUR (PERCENT RETENTION)
100
Figure 9-4. Distributions of measured and modelled sulfate concentration and percent
sulfur retention for the M-APP target population of 5,496 stream reaches.
Differences between this figure and plots C and D of Figure 9-3 reflect the
unequal weighting factors for individual sample watersheds used to generate
regional population estimates.
176
-------
Table 9-3. Summary of Sulfur Status in Surface Waters for Watersheds in DDRP Target
Populations in the NE, M-APP, and SBRP
Variable
SO4 concentration
(«eq/L)
Region
M-APP measure
model
NE measure
model
SBRP measure
model
Mean
148
171
110
128
37
44
Std.Dev
55
74
40
55
26
23
Min.
55
40
34
52
15
13
Median
153
165
105
118
24
41
Max.
313
365
214
281
119
107
Steady state S04 M-APP measure
concentration
(weq/L) NE
SBRP
measure
measure
246
118
122
72
47
23
140
50
86
215
111
120
405
239
203
Change in sulfate M-APP measure 98
to steady state
concentration NE measure 8
("eq/L)
SBRP measure 85
-39
-51
25
107
0
83
248
28
167
Percent sulfur
retention
M-APP calculated8
model
NE
calculated8
model
SBRP calculated3
model
37
30
4
-8
71
64
24
22
20
4
17
18
-21
-54
-19
17
20
44
26
0
7
78
68
74
84
69
0
87
88
8 Calculated retention, based on measured NSWS sulfate concentrations and on estimates of precipitation, runoff, and sulfur
deposition as described in Section 5.6.
177
-------
LAKE/STREAM SULFATE
STEADY STATE SULFATE
1.0
1.0
0.0
50
100 150 200 250
SULFATE (neq/L)
300
350
O
-i "-i r
50 100 150 -200 250 300
SULFATE (neq/L)
350 400 . 450
PERCENT SULFUR RETENTION
CHANGE IN SULFATE TO
STEADY STATE
1.0
-60
-40
—r
-20 0 20 40 60
PERCENT SULFUR RETENTION
100
O
-50
50 100 150
DELTA SULFATE Qieq/L)
200
250
Figure 9-5. Distributions of measured sulfate concentration, steady-state sulfate concen-
tration, percent sulfur retention, and change in sulfate concentration to sulfur
steady state in the NE, M-APP, and SBRP.
178
-------
Table 9-4. Comparison of Sulfate Concentrations and Sulfur Budget Status for DDRP
Watersheds in Subregions of the M-APP Region3
Subregion n
1D 7
2B 9
20 20
2Cab 7
2Cbb 5
2Ccb 8
Measured SO4
(aeq/L)
125
55 - 166
124
69 - 209
174
74-313
192
96 - 313
188
154 - 264
149
74 - 223
Steady State SO4
(weq/L)
228
194-313
288
207 - 405
224
140 - 299
268
185 - 299
266
189 - 278
183
140 - 243
Percent Sulfur Retention
44.0
17.9- 71.8
57.2
48.5 - 73.6
21.4
-21 .5 - 64.0
27.5
-8.6 - 65.0
17.3
4.7 - 28.8
18.6
-21.5-61.0
"Values are unweighted mean concentrations and ranges.
b2Ca = First four characters of Watershed ID = 2C28, 2C29, or 2C35
2Cb
2Cc<
2C41
2C46, 2C47, or 2C57
179
-------
The current status of surface water sulfate and of sulfur budgets in the M-APP suggests a region
that is transitional between NE watersheds, most of which are at or near steady state, and those in the
SBRP, all of which are still retaining sulfate. Although the budget status supports such a "transitional"
state, there is no clear north-to-south gradient of increasing sulfur retention (Table 9-4). The highest
average sulfur retention in the region (57 percent) occurs in stream survey subregion 2B (Valley and
Ridge Province in Pennsylvania, Maryland, West Virginia and Virginia), and the lowest average
retention (21 percent) occurs in the Allegheny Plateau (stream survey subregion 2C, in Pennsylvania,
Maryland, and West Virginia). Within the plateau region, sulfur retention is highly variable; 3 (of 20)
stream systems in the sample population have net retention higher than 40 percent, but an equal
number are computed as having net sulfur release. Within the same region, a trend of decreased
deposition from north to south is matched by proportional decreases in stream sulfate concentration,
and no spatial trend for retention is apparent. In the Valley and Ridge, there are similar trends of
decreasing deposition and stream sulfate from north to south, but no trend in percent retention.
9.2.4.4. Model Results -- Projections for Current Sulfur Deposition
The principal focus of the Level II analyses is to project the future dynamics of sulfate in the
M-APP watersheds. Figure 9-6 shows temporal trends for both the past and the future (for constant
future deposition) for soils in a typical M-APP watershed, as well as for the fastest and slowest
responding systems among the sample watersheds. As might be expected based on sulfur budget
data, model projections indicate a wide range of response times to sulfur deposition for soils in the
region. For many systems in the region, there is a lag of 25 to 50 years between increases in sulfur
deposition and corresponding sulfur outputs, resulting in current sulfur retention of between 30 and 40
percent. Some systems, however, are projected to respond with very short lags of only about 10 to
15 years, and are projected to be already at steady state. At the other extreme, a few watersheds with
deep, highly-adsorbing soils have very long lags and are projected to have only recently begun to
experience sulfate breakthrough from the soil. Some of the latter systems are not projected to reach
sulfate steady state for over 150 years.
Comparison of time traces for the M-APP watersheds to those for the NE and SBRP indicate
major differences among the three regions, in terms of both historic deposition and response time of
soils to changes in deposition. In the NE, significant increases in sulfur deposition began much earlier
than in the M-APP Region, almost a century ago, and deposition reached current levels around 1910.
During much of the intervening time, deposition exceeded current levels, at times by as much as 60
percent (Gschwandtner et al., 1985). Along with very high historic loadings of sulfate, soils in the
region have low adsorption capacities and respond with very short lags to changes in deposition.
Previous DDRP analyses for the region (Section 9.2 of Church et al., 1989) indicate that soils in the NE
respond over time frames of about a decade.
Simulations for watersheds in the SBRP used a historic input sequence (based on Gschwandtner
et al., 1985) in which sharp increases in deposition occurred later than in the M-APP. Modelled depo-
sition inputs increased almost exponentially from the late 1800s to 1970 (with a decrease during the
1930s and a temporary increase during the 1940s), but inputs did not exceed 50 percent of current
loadings until 1955 and have never significantly exceeded current levels of deposition. Sulfate
180
-------
SULFATE RESPONSE TIME
DDRP MID-APPALACHIAN WATERSHEDS
120
X
100-
I ou
LL
=! 601
CO
LJJ
> 40"
ULJ 20-
CC
0
Deposition Input
Stream Output
Typical
Range
T
1850
1900
1950' 2000 2050 2100 2150
YEAR
Figure 9-6. Time trends for modelled inputs and outputs of sulfate for soils on a typical
watershed in the M-APP Region, and for the fastest and slowest responding
systems in the region. Historic inputs are based on Gschwandtner et al.
(1985). Sulfur flux is expressed on a relative scale, with base year (1986)
deposition flux equal to 100. The base year is indicated by the arrow.
Because annual precipitation and runoff are constant during the simulation
period, input and output curves also simulate relative changes in sulfate
concentration. Simulated changes in sulfate concentration in precipitation are
proportional to changes in deposition flux, and changes in streamwater sulfate
to output flux (scales for concentration of precipitation and runoff are not
equal, due to dry deposition and evapoconcentration of sulfate in runoff).
181
-------
response times for SBRP watersheds, as for systems in the M-APP, are characterized by a wide range
of response times. Lags following changes in deposition range from about 20 to more than 150 years,
with an average of roughly 50 years. The differences in sulfur budget status between M-APP and
SBRP watersheds result both from differences in adsorption capacity and from the earlier onset of high
deposition inputs to M-APP watersheds.
Responses of M-APP watershed soils to a scenario of constant future deposition are summarized
by Figures 9-7 and 9-8 and in Table 9-5. Figure 9-7 shows distributions of modelled sulfate concentra-
tion and percent sulfur retention for the base year of 1986, then at 20, 50, and 100 years and at sulfur
steady state (concentration only); also shown are projected changes (increases) in sulfate concentra-
tion between the base year and the same 20-, 50-, 100-year, and steady-state intervals. Data for
simulations exceeding 50 years are included principally to indicate the relative magnitude of change
between 50 years and steady state, compared to projected changes from the present to year 50.
During the next 50 years, substantial increases in streamwater sulfate are projected, during which time
soils in a majority of watersheds in the region will approach sulfate steady state. During the next 20
years, median sulfate concentration is projected to increase by 25^eq/L (from 165 to 190/feq/L), with
a mean increase in concentration of 32/^eq/L (maximum increase of 76,weq/L). During the same inter-
val, percent sulfur retention is projected to decrease by an average of 15 percent, and the fraction of
M-APP watersheds with less than 5 percent sulfur retention is projected to increase from 14 to 36
percent. The projected increase in median sulfate concentration at year 50 is 41 ,«eq/L (165 to 206
/^eq/L) (projected change in median sulfate is not necessarily equal to the median projected change),
with an accompanying decrease in median retention from 26 to 3.1 percent. After 50 years, soils on a
majority (58 percent) of watersheds in the region are projected to be within 5 percent of sulfur steady
state, and retention is projected to exceed 10 percent in only 23 percent of M-APP watersheds.
For time periods longer than 50 years, the projected rate of change for sulfate in soils of
M-APP watersheds slows considerably, due to the high percentage of systems that have already
reached or are near steady state. Between years 50 and 100, median sulfate is projected to increase
by only 4^eq/L and median percent retention to decrease by less than 3 percent (from 3.1 to 0.4
percent). At year 100, only 7 percent of watersheds are projected to have more than 5 percent sulfur
retention. Those sites are characterized by soils formed from volcanic parent material that is more
similar to soils of the SBRP than to the soils of other M-APP watersheds that have formed from sedi-
mentary materials. Projected changes after 100 years are minimal, since most sites are projected to
be essentially at steady state by that time; increases in sulfate concentration between 100 years and
steady state are projected to exceed 1 yueq/L for only 42 percent of M-APP watersheds and to be
greater than 10 jueq/L in less than 10 percent of watersheds. Most of the sites projected to have
continued retention after 100 years are located in the Blue Ridge or Piedmont of Virginia.
Time to sulfur steady state for soils in M-APP watersheds, for a scenario of constant sulfur depo-
sition, is summarized in Figure 9-8. Consistent with procedures used in previous analyses for the NE
and SBRP and as discussed earlier, response time (time to sulfur steady state) has been defined as
the time for sulfate to reach 95 (or 105) percent of steady state concentration. In the M-APP Region,
soils on roughly 15 percent of the watersheds have already reached sulfur steady state. Median
projected time to steady state is 35 years; 25th and 75th percentiles of projected time to steady
182
-------
SULFATE CONCENTRATION
50 100 150 200 250 300 350 400 450
0.0
SULFATE (neq/L)
CHANGE IN SULFATE CONCENTRATION
TO STEADY STATE
o
80 120 160
SULFATE (neq/L)
200
240
PERCENT SULFUR RETENTION
T
40 80 120 160 200 240
PERCENT SULFUR RETENTION
Figure 9-7.
Projected changes in sulfate concentration, delta sulfate (change from base
year concentration), and percent sulfur retention for M-APP DDRP target
stream reaches.
183
-------
O
DC
O
Q_
O
DC
D_
LU
O
TIME TO SULFUR STEADY STATE
1.0
0.8-
0.6-
0.4-
0.2-
0.0
0 20 40 60 80 100 120 140 160 180
TIME TO 95% STEADY STATE (YEARS) AFTER 1986
Figure 9-8. Distribution of time to steady state for sulfur for soils on DDRP target stream
reach watersheds in the M-APP Region.
184
-------
Table 9-5. Summary Statistics for Modelled Changes in Sulfate Concentration and Percent
Sulfur Retention, and for Changes in Sulfate Concentrations for Watersheds in
the M-APP Region8
SO4_conc
%Srtnb
A_S04C
year to 95%
steady state
Year
measured
model-0
20
50
100
140
steady
measured
model-0
20
50
100
140
0-20
0-50
0-100
0-140
0-steady
—
A_SO4 - rate /zeq/LVyr
of change
in year 0
Mean
148
171
204
226
240
243
246
37 •
30
17
8 ,
3
1
32
55
68
72
74
46
2
Std dev
55
74
77
77
74
73
72
24
22
19
14
6
3
21
51
51
55
58
._
1
MIN
55
40
59
94
140
140
140
-21
-1
-0
0
0
0
-2
-3
^3
-3
-3
0
0
P_25
109
121
151
183
187
188
190
20
10 -
2
0
0
0
19
.35
38
38
38
6
. 2
MEDIAN
153 f
165
190
206
211
213
215
44
26
11
3
0
0
33
43
45
45
45
.35 -
2
P_75
175
208
237
268
283
287
293
56
45
25
9
2
1
39
75
99
101
103
62
4
MAX
313,
365
397
404
405
405
405
74
84
75
55
26
11
76
134
171
200
218
160
5
Model projections for future sulfur deposition at current levels.
% of systems with >5 % S retention
Year 0 86%
20 64
50 42
100 7.5
Change in modelled sulfate concentration from year 0 to n.
185
-------
state are 6 and 62 years, respectively, and the maximum projected response time is 160 years.
Between the base year and year 80, there is an essentially steady increase in the percentage of
systems at steady state, to about 90 percent of systems at steady state by year 80. Of those systems
with longer response times, most are represented in the 36 modelling watersheds by sites in the
Virginia Blue Ridge or Piedmont and have deep soils with very high adsorption capacities. Soils and
response times for these sites are similar to those previously characterized by DDRP analyses in the
SBRP.
The distribution of sulfate response times in the M-APP is°intermediate with respect to the
distribution in the NE and SBRP (Table 9-5, Figure 9-9). Previous analyses for the other regions,
discussed in Section 9.2 of Church et al. (1989), found that almost 40 percent of NE systems were at
steady state in the base year, and all systems were projected to be at steady state within 9 years
following the base year. In contrast, no SBRP watersheds are presently at sulfur steady state, and
median projected time to steady state is 61 years, almost twice the 35 year median for the M-APP
Region. The largest differences in response time between M-APP and SBRP stream systems lie with
those sites responding in the shortest period of time. In contrast to the 15 percent of M-APP sites
already at steady state, no SBRP sites are presently at steady state; the first SBRP watersheds are not
projected to reach steady state for 16 years. Similarly, 25 percent of M-APP sites are projected to be
at steady state within 6 years, compared to a projected 40-year period before as many as 25 percent
of SBRP sites come to steady state.
9.2.4.5. Model Results - Projections for Scenarios of Reduced Future Deposition
In addition to projections of future sulfur dynamics at constant deposition, simulations were run
for the two scenarios of reduced sulfur deposition show in Figure 9-1. Figure 9-10 compares the
response of a typical M-APP watershed and of fast- and slow-responding systems in the region to con-
stant deposition and to the two scenarios of reduced sulfur deposition. Probably the most important
result shown by these projections is that under conditions of reduced deposition, systems will even-
tually come to steady state at lower sulfate concentrations than they will if current levels of deposition
are maintained. The projections reported here are based on an assumption that sulfate adsorption by
soils is fully reversible. If sorption is only partially reversible, response times to decreases in
deposition will be shorter, because only a portion of sulfate will be desorbed. For individual systems,
the nature and timing of the response to decreased deposition are quite variable. For watersheds with
soils having low adsorption capacities that have responded quickly to historic deposition and are near
steady state for current deposition, decreases in deposition are projected to lead to rapid decreases in
sulfate efflux and to proportional decreases in sulfate concentration in runoff. Conversely, for the
systems that respond very slowly and that are still at less than 50 percent of steady state, a decrease
in deposition is projected to result in more gradual increases in sulfate as systems slowly approach
steady state with the lower level of deposition. For most watersheds in the region, a lagged response
comparable to that of the "typical" watershed in Figure 9-10 is projected, that is a continued increase
in sulfate export for one to three decades, as sulfate currently adsorbed on upper soil horizons, in
equilibrium with today's high levels of deposition, is desorbed and leached from the soil, followed by a
decrease in sulfate as the system comes to steady state with the lower level of deposition. The extent
and duration of such lags (i.e., the time interval between a decrease in deposition and the onset of
186
-------
TIME TO SULFUR STEADY STATE
NE, MID-APR, AND SBRP
Z)
o
1.0
o
p 0.8
DC
O
O 0.6
Q_
LLJ
> 0.4 H
0.2-
0.0
NE
M-.
0
—T—
40
—r—
60
—i—
100
—I—
120
—i—
160
20 40 60 80 100 120 140 160 180
TIME TO 95% STEADY STATE (YEARS)
200
Figure 9-9. Distributions of time to steady state for sulfur in soils on DDRP target
watersheds in the NE, M-APP, and SBRP.
187
-------
SCENARIO A
120
* 100
11.
L1I
80-
:D 60-
CO
ai
> 40-
LLI 20-
cc
1 1 1 1—
1850 1900 1950 2000 2050
YEAR
SCENARIO B
120
100-
LU
80-
r> 60"
00
LU
> 40-
UJ 20-
GC
1850 1900 1950 2000 2050
YEAR
SCENARIO C
120
100
u_
UJ
80
3 60
CO
HI
> 40-
Uj 20-
CC
1
2100 2150
2100 2150
1850 1900 1950 2000 2050 2100 2150
YEAR
Figure 9-10. Projected temporal trends for sulfate in a typical M-APP watershed, and for
the fastest and slowest responding systems in the region, for three scenarios
of future sulfur deposition. Sulfur flux is expressed on a relative scale, with
base year (1986) deposition equal to 100. The base year is indicated by the
arrow. Note that because annual precipitation and runoff are constant
throughout the simulation period, changes in sulfate concentration are
proportional to changes in sulfur flux.
188
-------
decline in projected streamwater sulfate concentration) in soils of M-APP watersheds is quite variable;
in simulations of the 36 M-APP watersheds (for Scenario B), the lag phase for some watersheds lasted
more than 40 years, during which time projected sulfate concentrations increased by as much as 60
fieq/L (Figure 9-11).
In addition to qualitative comparisons of sulfate concentrations at steady state, it is important to
compare scenarios of future sulfate to current sulfate concentrations. As shown in Table 9-6, at
current deposition, mean sulfate concentration is projected to rise sharply during the next 50 years
(171 to 226 fieq/L), with a subsequent increase to 240 ^eq/L in 100 years and to 246 fieq/L at steady
state. In contrast, under deposition scenario "B" (50 percent decrease in deposition between years 5
and 15, then constant at 50 percent of current deposition), simulations project a temporary increase in
sulfate, as described above, but then a decrease to 150, 133, and 129 /ueq/L at 50 and 100 years and
at steady state, respectively. For scenario "C", as described earlier, projected concentrations initially
rise, but decrease from 171 to les^eq/L by year 50 before increasing by about 20 percent to 191 and
199 /teq/L at 100 years and at steady state.
These results show that in the long run, sulfate outputs from soils and sulfate concentrations in
solution will be proportional to deposition. In the short term, sulfate concentrations will increase,
sometimes by a substantial amount, even if deposition is sharply reduced; within 50 years, however, a
reduction in deposition will result in substantial decreases in sulfate concentration below current
concentrations and especially in comparison to the concentrations that will result if deposition remains
at current levels, In order to maintain long-term sulfate concentrations at current levels, deposition
would need to be reduced to 63 percent of current input (i,.e., equal to average current output); higher
deposition would result in long-term increases in deposition.
9.2.5. Summary
The projected sulfate response of soils and surface waters in the M-APP describes a region
composed principally of delayed response systems. Although a small percentage of watersheds in the
region have soils with very short response times to sulfur deposition, most M-APP watersheds are
mantled with soils that have buffered sulfur inputs for several decades and are continuing, albeit with
diminishing efficiency, to retain sulfate by adsorption. At the present time, median modelled retention
in M-APP watersheds is 26 percent, but modelled sulfate concentrations are increasing as sulfate
breaks through the soil column, and retention is projected to decrease sharply over the next few
decades. Median modelled retention (assuming continued sulfur deposition at current levels) is
projected to decrease from 26 percent at present to 11 percent within 20 years and to only 3 percent
within 50 years. Sulfate adsorption has played an important role in minimizing the adverse effects of
acidic deposition on surface waters in many M-APP watersheds, but as sulfate concentrations increase
in the next few decades, there will be a proportional increase in sulfate-mediated cation leaching, due
to a combination of the salt effect and depletion of exchangeable base cations; effects of sulfate
leaching on cation removal from M-APP soils are discussed in Section 9.3 and in Section 10.
The nature of future sulfate dynamics in the soils of M-APP watersheds, as discussed in Sections
9.3. and 9.4, is highly dependent on the nature of sulfur deposition to watersheds of the region. At
189
-------
MID-APPALACHIAN WATERSHEDS
PROJECTED LAGS AFTER DECREASE IN DEPOSITION
^
Q)
3;
LLJ
I'-
LL
CO
LLJ
CO
LLJ
DC
O
80 -
70 -
•
60 -
*
50 -
40 -
30 -
20 -
10 -
o-
•
• •
B
B •
*
• " * • "
•••
"" B ."
m*
• •
_-|Q -| 1 1 1 1 1 1 1 1 ' 1 ' 1 • r i • i
0 5 10 15 20 25 30 35 40 45 5<
DURATION OF LAG (years)
Figure 9-11. Extent and duration of lags in modelled sulfate response for soils on 36
M-APP watersheds following a 50 percent decrease in sulfur deposition.
190
-------
Table 9-6. Summary Comparison of Modelled Changes in Sulfate Concentration in M-APP
Stream Systems Under Three Future Sulfur Deposition Scenarios8
Variable
Year
Scenario13
B
Measured SO4
(aeq/L)
148
148
148
Modelled SO4C
(weq/L)
AS04e
(aeq/L)
0
Iagd
50
100
steady
50
100
steady
171
~
226
240
246
55
68
74
171
200
150
133
129
-21
-38
-42
171
168
191
199
-A
19
28
a Concentrations are weighted means for the DDRP target population.
b Deposition scenarios as described in Section 9.2.3.2 and shown in Figure 9.1; A is constant future
deposition at current levels for the entire simulation period; B is constant at current levels for
5 years, decreasing to 50 percent of current levels by year 15, and constant, thereafter; C is constant at
current levels for 5 years, decreasing to 50 percent of current loadings by year 15, then increasing to 80 percent
of current levels si year 50 and constant thereafter.
0 steady «• steady state.
d Lag is maximum concentration during lag phase after reduction in deposition.
8 A SO4 = change in sulfate from year 0 to n.
191
-------
constant deposition, average sulfate concentrations are projected to increase by over 40 ^eq/L in the
next 50 years, whereas if deposition is reduced by 50 percent, the average concentration is projected
to decrease by 25 //eq/L. The projected difference in average concentration between the two deposi-
tion scenarios increases to almost 120 /leq/L at sulfur steady state. It is also important to recognize
that despite the lags in sulfate response that are projected following a decrease in deposition, pro-
jected sulfate concentrations for the scenario of decreased deposition are in all cases, at all times in
the future, lower than those projected if deposition is not reduced. It is also important to reiterate that
the projections and time frames discussed here refer only to sulfur dynamics as affected by sorption in
soils. Sulfate mobility is considered to be one of the principal factors influencing changes in transport
of cations from the soil, but processes that do not involve sulfur also play crucial roles in the dynamics
of base cations and ANC. Rates of change in ANC, and especially potential time to zero ANC, are not
necessarily coincident with time to steady state for sulfur. Systems can reach zero ANC before, con-
currently with, or after time to sulfur steady state. The relationship between changes in sulfate and
ANC will be characterized and discussed as part of the integrated watershed modelling (Level III)
analyses in Section 10.
The modelling projections described here are generally consistent with other sulfur analyses for
the M-APP in this report. Measured sulfate concentration and percent sulfur retention data, described
in the sulfur budget analyses in Section 7, provided part of the "true" data used for model calibration
and are therefore similar as a result of fitting during calibration of model inputs. Of greater signifi-
cance, the projections of future sulfate mobility are very similar to those generated as part of the
integrated modelling described in Section 10. Although the two models are based on the same sulfate
adsorption algorithm and on the same isotherm and deposition data, several important model proper-
ties (hydrologic routing and consideration of deep (> 2 m) soils) are different, and calibration
procedures were different and were conducted independently. Despite these important differences,
significant similarities exist between the two sets of projections for sulfate. These similarities contribute
to confidence in both analyses.
Finally, major results for the Level II sulfate analyses for M-APP watersheds can be summarized
as follows:
• Measured and modelled sulfate concentrations in M-APP watersheds are, on average, higher
than equivalent values for NE and SBRP watersheds. Because sulfur deposition to M-APP
watersheds is substantially higher than to most NE or SBRP watersheds, most M-APP
systems are still retaining sulfur.
• Median measured sulfur retention in M-APP watersheds is 44 percent; median modelled
retention is somewhat lower at 26 percent; modelled retention ranges from -1 to 84 percent.
• If current levels of sulfur deposition are maintained, sulfate concentrations are projected to
increase substantially over the next 50 years; the projected median increase is 33 ^aeq/L over
the next 20 years and 43 /teq/L in the next 50 years. Maximum projected increases for the
same intervals are 76 and 134/ieq/L. Increases in sulfate concentration between year 50
and steady state are projected to be much smaller than during years 0 to 50, because most
systems will be near steady state within 50 years.
192
-------
• As sulfate concentrations increase, median percent retention is projected to drop from 26
percent at present to 11, 3, and 0.4 percent at 20, 50, and 100 years, respectively. For the
same time intervals, the projected fraction of the target population at sulfur steady state (i.e.,
< 5 percent retention) is projected to increase from 14 percent in the base year to 36, 58,
and 93 percent.
* If sulfur deposition is reduced, there will be proportional decreases in projected steady-state
sulfate concentration. Intermediate dynamics are dependent on the amount and timing of
decreases in deposition.
• For systems with sulfate concentrations significantly lower than steady state, there is likely to
be a lag phase following any major reduction in deposition, during which sulfate concentra-
tions in soil leachate continue to increase due to desorption of previously adsorbed sulfate.
In projections for M-APP systems, such lags last as long as 50 years, during which time
sulfate concentrations rise as much as 60 ^eq/L During any such lag phase, sulfate con-
centrations would be lower than if deposition had not been reduced.
• Qualitatively, sulfate adsorption by soils has been, and continues to be, an important
process in delaying the effects of acidic deposition in M-APP watersheds. Most M-APP
systems should be regarded as delayed response systems with regard to sulfate. The
relative importance of adsorption is much higher in M-APP systems than in soils of NE
watersheds. It is somewhat lower than in soils of SBRP watersheds, however, due both to
the lower average potential for adsorption by typical M-APP watersheds and to the very low
adsorption (and short response times) for a small percentage of M-APP watershed soils.
9.3 EFFECT OF CATION EXCHANGE AND WEATHERING ON SYSTEM RESPONSE
9.3.1 Introduction
Soils provide an important linkage between acidic deposition and surface waters. Soil can
buffer the effects of acidic deposition by a variety of processes including cation exchange, mobile
anion absorption, and mineral weathering. The potential of soils to buffer acidic deposition is often
quite large, especially in the case of calcareous soils. In the case of M-APP uplands, however, soils
have formed from sedimentary bedrock that consists of the slowly weatherable remnants of previous
weathering cycles. These rocks and associated soils often contain very low quantities of base cations,
and thus have a limited capacity to neutralize acid inputs. Consequently, not only are surface waters
in this region at risk from acidic deposition, the soils in the region are already acidic and could
become even more acidic.
During the development of the NAS panel report (MAS, 1984), much discussion was devoted to
the role of cation exchange and mineral weathering in protecting watersheds from acidification. One
group of panel members argued that soil cation exchange in most watersheds provides a large reser-
voir of base cations, sufficient to buffer against potential changes caused by acidic deposition.
Therefore, they argued, even in watersheds where cation exchange (as opposed to mineral weather-
ing) is an important process in supplying base cations, the future effects of acidic deposition are
probably not a concern. Another group of panel members argued that the buffering capacity of soils
was finite, and that continued exposure to current levels of acidic deposition would have long-term,
adverse effects on water quality in some systems. The conclusion of the committee as a whole was
that the role of cation exchange in buffering against the effects of acidic deposition is an area of
193
-------
considerable uncertainty, and that these processes need to be considered when attempting to project
future effects of acidic deposition on aquatic ecosystems. Thus, the Level II base cation studies were
designed to determine the role of base cation exchange in controlling future changes in surface water
chemical composition. The specific objectives were to:
• identify the role that base cation exchange has in determining current surface water
composition;
• determine the capacity of base cation exchange processes to buffer against future changes
in surface water composition as a result of acidic deposition; and
• make projections regarding the magnitude and extent of changes that could occur in region-
ally representative soils and surface waters as a result of continued exposure to acidic
deposition.
We developed several hypotheses to guide the investigations of the role of cation supply
.processes in regulating surface water chemistry in representative watersheds in the DDRP study
regions.
9.3.1.1 Level II Hypotheses
Five hypotheses guide the Level II base cation analyses:
• Cation exchange processes determine surface water composition.
• Soils delay surface water acidification.
• Increased deposition induces net cation leaching.
• Cation resupply rate is slow.
• Soil chemistry is an indicator of soil response to acidic deposition.
A brief discussion of each hypothesis follows.
9.3.1.1.1 Cation exchange processes determine surface water composition --
The first hypothesis is designed to identify the primary process or processes that regulate surface
water chemistry. In systems that have attained steady state with respect to sulfate deposition (see
Sections 7 and 9.2), primary mineral weathering and biological uptake are probably the principal
processes that modify the composition of incident deposition. Under steady-state conditions, the base
cation exchange pool should actively reflect the dynamic balance between these two important proc-
esses. Regardless of their relative importance, however, if soils regulate surface water ANC values,
then this should be reflected by the composition and chemical properties of the soil exchange complex.
The hypothesis is tested by comparing surface water composition projected using soil cation
exchange models with observed values. A close correspondence between the observed and projected
values suggests that soil exchange processes have a major role in regulating surface water chemistry.
194
-------
9.3.1.1.2 Soils delay surface water acidification --
The second major hypothesis is that soils will delay, but not prevent, the acidification of surface
waters. The concept behind this hypothesis is that soils have a finite capacity for buffering against
changes in surface water chemistry caused by increased levels of acidic deposition. In essence, the
chemical and physical characteristics of a soil reflect its response to some given set of environmental
conditions. Therefore, at a given level of deposition, vegetative-uptake, mineral weathering, etc., the
cation exchange pool reflects a balance of the various sources and sinks for cations in that area. This
balance is dynamic, changing with the seasons and with the shifting flow of cations among the various
reservoirs.
When a perturbation such as acidic deposition is imposed on a system, the system (in this case
the soil) evolves toward a hew state of balance. The rate at which changes take place depends both
on the sizes of the cation reservoirs in the system and on the flux of material between reservoirs. If
the transfer rates of material between reservoirs is slow, or if the mass of material in the affected
reservoirs is large relative to the transfer rates, then rates of evolution toward the new balance point
tend to be slow. Conversely, if reservoirs are small relative to the size of the flux between reservoirs,
then adjustments to a new system steady state can occur rapidly.
This hypothesis contends that the pool of base cations on soil exchange complexes is large
relative to the rate of cation loss from the system by leaching. As a result, the rate of adjustment of
the exchange complex to the new deposition conditions should require decades to centuries before a
new steady-state, or dynamic balance, condition is attained.
The hypothesis is tested using a model approach. Measured soil properties serve as inputs to
the various models. The Level II models, all of which have a mass balance component, track the loss
of base cations from soils at the specified levels of deposition. The models are concerned primarily
with exchange processes and do not explicitly include cation supply via weathering. Therefore, the
computed mass balances should correspond to the maximum leaching rates that could occur. The
rate of change of base cation status of the soils included in the study, then, should be related to the
amount of time over which the soils should delay acidification of surface waters.
9.3.1.1.3 Increased deposition induces net cation leaching --
The third major hypothesis is that increased levels of deposition, specifically, increased concen-
trations of sulfate and nitrate in deposition, increase the rate of cation leaching from the soil exchange
complex by way of the mobile anion effect (Johnson et al., 1980; Seip, 1980). Two factors are consid-
ered in evaluating this hypothesis. First, the average base status of the soil exchange complex repre-
sents a balance among the various supply and demand processes in the ecosystem. For example,
under steady-state conditions, weathering should supply sufficient base cations to meet the demands
of vegetative uptake while maintaining soil solution concentrations in equilibrium with the soil
exchange complex. Perturbations to the system, such as changed deposition, will alter this steady-
state condition.
195
-------
Second, charge balance requirements need to be maintained between the soil exchange com-
plex and soil solutions. Maintaining charge balance, coupled with the increased anion loads provided
by acidic deposition, requires that total (acid plus base) cation concentrations in soil solutions
increase. The ratio of the base to acid cations will not change dramatically, however, at least during
the initial stages of leaching. The higher concentrations of base cations in soil solutions lead to a net
depletion of base cations from the exchange complex. If this increased leaching is not matched by an
increased level of supply (e.g., from weathering), then the overall effect will be a net depletion of the
base cations from the exchange complex.
9.3.1.1.4 Cation resupply is slow —
The fourth hypothesis, that the rates of cation resupply to the soil exchange complex are slow
relative to the rates of base cation removal, is not being tested directly in this study. Rather, the
hypothesis is being subsumed in the models as an assumption that exchange reactions provide suffi-
cient buffering for resupply rates not to be an issue for the time scales of concern to the study. As
such, mineral weathering rates have been defined as zero for these Level II simulations. Simulations
that include batch weathering and exchange processes are presented and discussed in Section 10.
One reason for using this approach is that, with current technology, no definitive method exists
for distinguishing the different sources of base cations to surface waters. Therefore, by assuming that
all base cations are derived from exchange sites, simulations yield, effectively, a "worst case" scenario
for the depletion of the soil buffering capacity. If, under these circumstances, the results suggest an
extensive capacity of the soil to buffer against the effects of acidic deposition, then the resupply rate is
not an issue of importance in this study.
9.3.1.1.5 Soil chemistry as an indicator of soil response to acidic deposition -
The final hypothesis is intended to provide the groundwork for using selected soil properties as
qualitative indices of soil "condition" and expected soil response to acidic deposition. Recently,
several attempts have been made to correlate soil properties with the anticipated response of soils to
acidic deposition (VanLoon, 1984; Stuanes, 1984; Lau and Mainwaring, 1985). Results from these
studies suggest that soil properties are useful indicators of how soils will respond to continued
exposure to present or anticipated levels of acidic deposition.
9.3.1.2 Approach
As discussed in Section 9.1, the approach used for Level II base cation analyses is model
based. The primary processes believed to regulate exchange processes are known, and models have
been developed that describe these processes with internal consistency. As such, existing model
formulations are used extensively in conducting these studies.
Data used in running the models were collected specifically for this study. In collating the data
for use in the models, we made certain decisions regarding how data from individual soils and
watersheds would be condensed, or aggregated, for use in the models. Because the primary goal of
196
-------
the DDRP is to make regionally representative projections about future changes in surface water
chemistry as a direct result of acidic deposition, we decided to aggregate the data, first into groups of
soil with similar chemical and physical characteristics, and then to the watershed level. The models
were run on watershed level data. The results from these runs were then regionally weighted so that
inferences about the DDRP target population of systems could be made.
9.3.1.2.1 Off-the-shelf models --
In designing the Level II base cation studies, one of the issues considered was the selection of
models. We decided during the planning stages of DDRP to use only published, peer-reviewed, and
publicly available models. A primary advantage of this decision was that the data requirements for
these models were known, thus field sampling programs could be developed to collect the appropriate
data. A second advantage was that minor modifications or improvements could be incorporated into
the model codes in a timely manner. The selected models have been developed and described in
detail by Bloom and Grigal (1985), Reuss (1983), and Reuss and Johnson (1985).
9.3.1.2.2 Aggregated soil chemistry data —
Following selection of models for use in the Level II analyses, a major issue was preparation of
data for use in the models. Soil physical and chemical data were gathered for a regionally represen-
tative sampling of soils in the NE and SBRP. These data were obtained from individual pedons and
soil horizons. V/e transformed these data into a form usable by the models, by aggregating the data
to produce information representative of whole watersheds.
Details of the aggregation procedures have been presented by Johnson et al: (1990). Briefly,
the steps taken to produce the aggregated data depend on the structure of the model to be applied.
In general, data are first averaged within the master horizons (i.e., O, A/E, B, or C horizons) of
individual soil sampling classes. Then, if required by the models (e.g., those that describe the soil as
a single "box"), results from the master horizons are averaged to yield a single value for each
parameter, which represents the sampling class as a whole. The procedures used to average soil
chemical and physical properties at the horizon and sampling class levels varied slightly in accor-
dance with the model for which the data were being developed. For models that use capacity varia-
bles as inputs (e.g., the Bloom-Grigal model), soil properties were averaged using mass weighting
procedures. For models using intensity variables as inputs, an intensity weighting scheme (Johnson
et al., 1990) was developed that preferentially weighted the lowest subhorizon in generating values for
master horizons, and then employed straight numerical averages to produce sample class/pedon data.
Finally, data from individual sampling classes were weighted, using areal weighting, to produce
soils data representative of the watershed as a whole. The weighting used in this last aggregation
step was proportional to the relative areal occurrence of the sampling class on a particular watershed.
The weighting effectively assumes that long-term changes in runoff chemistry are controlled by the
extent of soil occurrence on a watershed, not by the location of soils on that watershed. For example,
although it might be argued that riparian soils have a greater influence on the composition of surface
waters than do ridge-crest soils, riparian zone soils and those soils immediately adjacent to the lakes
197
-------
are not preferentially weighted relative to upland soils. The decision to use the uniform weighting
approach was based primarily on the difficulty of developing uniform, broadly based algorithms to
apply preferential weighting to specific soils based on geomorphic considerations.
9.3.1.2.3 Scale of temporal forecasts-
Another decision to be made in implementing the Level II analyses was the time scale over
which the model simulations would be run. In the near term, dramatic, permanent changes to surface
water composition are not expected to occur on annual time scales. Acidic deposition has affected
parts of eastern North America for at least several decades. As such, changes in soil and runoff have
probably already occurred in many soils and surface waters; some rapidly changing systems might
have already changed substantially and are near steady state with current high levels of acidic
deposition, whereas soils with relatively large base cation pools might be in incipient stages of
response, and would be expected to change only slowly.
For long time-scale projections, the major factor determining the duration of simulations to be
run is the uncertainty associated with the major parts of the modelling efforts. As soil composition and
properties evolve with continued exposure to acidic deposition, the response of these soils is also
expected to change. We anticipate that, for longer time scales, projected changes will become more
dramatic. The larger changes are balanced, however, by increases in the uncertainty of the analyses
for periods exceeding, e.g., 50 years.
Using these procedures as guides for bounding the time intervals to be modelled, model
simulations were run for 100 years. The results are presented as projected changes after 20, 50, and
100 years. The 20- and 50-year projections provide information about relatively near-term changes
that might be anticipated and are relevant time frames with regard to the potential implementation of
regulatory controls.
9.3.2 The Bloom-Grigal Mode!
The Bloom-Grigal soil acidification model (Bloom and Grigal, 1985) is a simple mass action
(nonequilibrium) exchange reaction model. The amount of acidity in deposition displaces an equiva-
lent amount of base cations and/or aluminum from the soil. The model then assumes that these base
cations are completely mobile and are removed from the system, causing a decrease in the overall
base saturation of the soil. Bloom and Grigal (1985) use a semi-empirical equation to relate soil base
saturation to soil pH. Soil pH is then used to determine the activity of AI3+ in soil solution. The
purpose of this analysis is to make a model-based analysis of the effects of the potential long-term
acidic deposition on the soils in the M-APP Region and to extrapolate these results to make regional
projections.
9.3.2.1 Model Formulation
The Bloom-Grigal model is formulated around the assumption that, in steady-state ecosystems,
acidic deposition depletes base cations from the soil ion exchange complex. The simplicity of the
198
-------
model lies in the fact that soils are treated as a single homogeneous unit or compartment and all
incoming deposition reacts completely with the soil in that compartment. The simplicity of the Bloom-
Grigal model makes it a useful tool for assessing the effects of acidic deposition on forested soils for a
variety of deposition scenarios and for long periods of time.
The Bloorn-Grigal model assumes that acidity in deposition reacts completely with the soil. In
other words, the model makes no provision for deposition to be routed around the soil and directly
into the surface water or into the subsoil strata. The amount of exchangeable base cations removed
from the soil compartment is calculated as the difference between the input acidity and the output of
H+ and A!3+, corrected for the protonation of bicarbonate. The amount of base cations lost is sub-
tracted from the pool of exchangeable base cations and a new base saturation is calculated. The
BIoom-Grigal model then calculates a new soil pH based on an equation that relates soil pH to base
saturation. After adjusting parameters, the model then simulates the next year of deposition (Figure
9-12).
The Bloorn-Grigal model estimates the loss of base cations on an annual basis using the
following equation:
S = I - A - C (9-3)
where S is the sum of base cations, I is the amount of effective acidity in deposition, A is the acid
leached from the soil, and C is the correction factor for the decrease in acidity due to protonation of
bicarbonate.
This model was created to assess the effect of acidic deposition on soils that do not adsorb
sulfate. Soils that adsorb sulfate have lower anion flux through the soil, and thus lower base cation
removal rates (at a given level of deposition) than nonadsorbing soils. In this regard, the BIoom-Grigal
model will probably overestimate the rates of base cation removal from the soils in the M-APP Region,
because many soils of the region are adsorbing and retaining sulfur. Retention in watersheds in the
M-APP ranges from -21 to 74 percent, with median retention of 44 percent (Section 7.9.2).
The BIoom-Grigal model incorporates the input of nitrogen in deposition. Because forested soils
are generally deficient in available nitrogen, inorganic nitrogen in deposition is removed by plants and
organisms in the soil (Bloom and Grigal, 1985). When plants assimilate nitrogen in the form of nitrate
(NO3"), they release hydroxyls (OH") to the soil, which is a nonacidifying reaction. When plants assimi-
late nitrogen as ammonium (NH4+), however, they release protons (H+), thus ammonium uptake is an
acidifying reaction. The biological oxidation of NH4+ to NO3' produces two H+ ions for every mole-
cule of NH4+ oxidized. The BIoom-Grigal model incorporates these processes in calculating the net or
effective acidity of deposition.
The original versions of the BIoom-Grigal'model were coded in FORTRAN and BASIC; the
version used in this analysis is coded in a high speed, compilable form of BASIC. In addition, code
has been added that allows the input data to be processed in a batch mode. The output is formatted
to magnetic media to simplify the data reduction process. Code also has been added to adjust
199
-------
'•• 'i', ss; ;
v- . --•;- - --',--' -'.
Record watershed
soil pH, base saturation
Calculate protonation
of bicarbonate
Calculate loss of bases
Calculate new base
saturation, soil pH and
soil solution AI3+
Deposition
Precip/runoff
Soil chemistry
Modelling
dataset
Figure 9-12. Flow diagram for the one-box Bloom-Grigal soil simulation model.
200
-------
deposition for the nonconstant deposition scenarios as the model is running. The fundamental
equations in the original model have not been altered, however.
9.3.2.2 Assumptions
We make a number of assumptions in modelling the effect of acidic deposition on soils with the
Bloom-Grigal model. Some are implicit to the model, others meet the needs of our current applica-
tion. The assumptions used in implementation of the Bloom-Grigal model are itemized below with
explanatory discussion or comments.
9.3.2.2.1 Sulfate adsorption --
The Bloorn-Grigal model assumes that sulfate is not adsorbed by the soil and acts as a com-
pletely mobile anion. As mentioned previously, in soils that have net sulfate adsorption, this
assumption may lead to an overestimation of the amount of base cations actually leached from the
soil. If the sulfate adsorption/leaching information from the preceding section were integrated into or
with the Bloom-Grigal model, the effect of sulfate adsorption on base cation leaching could be
assessed. These two analyses are not coupled in the Level II analyses, but comparable integration is
used with the Level III integrated watershed modelling analyses (Section 10). Consequently, the
analysis presented here represents a worst-case scenario of base cation removal from soils as a result
of acidic deposition.
9.3.2.2.2 Input acidity--
The total effective acidity (H+tota|) in deposition was defined by Bloom and Grigal, and is
similarly computed in our analyses, as:
(9-4)
Briefly, forested ecosystems are generally nitrogen limited. Therefore, nitrogen deposition in these
systems can contribute to the total system proton budget (see Section 9.3.2.1 above). Equation 9-4
takes this into account. Bloom and Grigal (1985) provide a complete discussion of the rationale for
this definition of input acidity.
9.3.2.2.3 Extent of reaction -
For this analysis, it is assumed that the input acidity, the effective acidity, reacts completely with
the soil. The acidity is partitioned between solubilizing Al and replacing exchangeable base cations.
This represents a worst-case scenario of base cation removal from soils as a result of acidic deposition.
9.3.2.2.4 Depth of soil --
The depth of reactive soil material equals the mean aggregated thickness of the soil sampling
classes represented by the types of soils on the specific watersheds. In their original paper, Bloom
201
-------
and Grigal (1985) considered the effects of deposition only within the rooting zone of soils (the top
25 cm). We consider the effect, however, on the whole soil compartment. Our soil chemistry input
data are aggregated to represent the central tendency of the soil chemical characteristics of the whole
soil compartment. The effect of acidic inputs on data aggregated in this way thus represents a mean
effect. Actual effects are likely to be greater in some parts of the soil profile than in others.
9.3.2.2.5 Volume of drainage water -
The volume of water moving though the watersheds in each year of simulation is equal to the
long-term annual average runoff. Other than the assumption that water moves vertically in this
analysis, no other assumptions are made regarding watershed hydrologic flowpaths.
9.3.2.2.6 Partial pressure of soil CO2 -
The partial pressure of ambient CO2 is approximately 0.0003 atm. Soil air is, however, enriched
with CO2 due to biological respiration and is consequently elevated. In all of the Bloom-Grigal model
runs, the partial pressure of CO2 in the soil air is set at 0.005 atm, a value thought to be reasonable
for forested soils (Fernandez and Kosian, 1987).
9.3.2.2.7 Activity of soil solution AI3+ »
To calculate the amount of input acidity that is converted to output acidity by aluminum
buffering, the activity of Al in soil solutions is calculated using the following equation:
log(A!3+) = 2.60 - 1.66 * soil pH
(9-5)
This equation in the Bloom-Grigal model represents an empirical relationship based on soil chemical
data for the Upper Midwest. In developing their model, Bloom and Grigal (1985) had a fundamental
,3+
problem in using the solubility of AI(OH)3 to describe the variation in Ar+ with pH. They state that in
very acidic soils, such as forested soils, Al is undersaturated with respect to the solubility of AI(OH)3.
Therefore, AI(OH)3 solubility was regarded as a poor model for the pH-A!3+ relationship. To establish
a more realistic relationship between AI3+ and pH, they used laboratory measurements of AI3+ in arti-
ficially acidified soils to define an empirical relationship between pH and Al3+.
9.3.2.2.8 Relating soil solution pH to base saturation --
The pH of soil solutions is related to base saturation (BS) by the following equation:
pH = pKa + n * log [BS/(1 - BS)]
(9-6)
where pKa is the apparent acidity constant for soil (i.e., aggregate watershed/soil compartment) and n
is an empirical constant. This equation is an extended form of the Henderson-Hasselbach equation.
202
-------
The Bloom-Grigal model used here calculates pKa and n for each watershed using the initial
aggregated values of soil pH and base saturation. These parameters describe the relationship
between soil pH and base saturation and are unique for each watershed.
9.3.2.2.9 Base cation uptake ~
The model assumes no net accretion of base cations in biomass. The uptake of base cations
by forest vegetation is an acidifying process by which H+ is exchanged for an equivalent amount of
base cations to maintain charge neutrality. At the same time, through litterfall and decomposition,
base cations are released to soils. The Bloom-Grigal model tracks only the flux of base cations
leached from the soil. This no-accretion assumption implies that the uptake of base cations by vege-
tation is exactly equal to the amount recycled to the soil (i.e., steady state). The model does not take
into account the potential loss of base cations by erosion or chemical complexation followed by leach-
ing reactions. Because most of the forested systems considered by the DDRP are second-growth
systems still in an aggrading phase, vegetation is probably a significant cation sink in most water-
sheds, and the assumption of no net accretion underestimates potential depletion of soil base cations.
9.3.2.2.10 Mineral weathering —
Mineral weathering is the major ultimate source of base cations to soils and runoff. The original
Bloom-Grigal model code had a subroutine that calculated the contribution of base cations to the soil
solution via mineral weathering. The rate of mineral weathering for the simulations presented here is
set to zero for three reasons. First, by assuming no base cation resupply, a "worst-case" base cation
loss scenario is evaluated, thereby bounding the projections. Second, the relationships between
weathering and soil solution pH are not sufficiently established to provide accurate parameters for the
weathering equations. One complication, in particular, is that mineral weathering rates are a dynamic
function of the chemical weathering environment. Third, the use of this subroutine of the Bloom-Grigal
model requires an initial estimate of the mineral weathering rate. These estimates are not available at
this time.
9.3.2.2.11 Cation exchange capacity -
The cation exchange capacity (CEC) of soils is held constant throughout the period of simula-
tion, although this approximation is not correct in soils if the soil pH changes. Soil CEC is derived
from two sources: (1) secondary clay minerals with permanent charge due to isomorphous substitution
of lower valent cations for cations in the clay crystal lattice, and (2) variable charge sites on organic
matter, para- and noncrystalline hydrous oxides, and edge sites on permanently charged clays. The
variable charge CEC is a function of pH (i.e., the net soil CEC changes as with changes in pH). As
pH increases, the variable charge CEC increases, and vice versa. Because of scientific and data
limitations, we have, however, chosen to hold CEC constant.
203
-------
9.3.2.2.12 Time steps —
The time step for simulations is annual. For assessment purposes, yearly time steps are a use-
ful increment. From a modelling standpoint, any shorter time step (e.g., daily) is data intensive and
computationally demanding. Shorter time steps may provide more accurate short-term projections
(e.g., interannual variability in soil chemistry), but for regional assessments, longer time steps appear
to be more useful and require less data.
9.3.2.3 Limitations
Soils are highly complex and no simulation models exist that accurately depict the fluxes of
energy and matter in soil systems. Models are based upon simplifying assumptions and are therefore
necessarily incomplete representations of reality. As with any attempt to project future events, the
Bloom-Grigal soil simulation model is not without limitations. Some of the limitations are due to the
state of the science and others are have been imposed by the DDRP.
The scientific limitations center around the factors that control aluminum solubility and the
relationship of soil pH to base saturation. Bloom and Grigal (1985) developed empirical equations to
describe this relationship for a selected set of forested soils in northeastern Minnesota. As described
in their paper, the equations appear to be appropriate for forested soils in Minnesota. For the DDRP,
the relationships described by Bloom and Grigal are assumed to be widely applicable and were not
independently verified. It is doubtful, however, that these or any other single equation would be
universally true due to spatial differences in soils and vegetation.
Soils are dynamic systems. Soil properties fluctuate on a daily basis, and daily temperature and
moisture changes affect a broad range of soil processes. Broader seasonal changes also occur. The
use of annual time steps assumes that soils are static, possibly restricting the accuracy of the pro-
jections. As mentioned in Section 9.3.2.2.12, however, shorter time steps are data intensive and
computationally restrictive.
Individual soil processes are inextricably linked to a number of other processes and any consid-
eration of a single process (e.g., base cation flux) in isolation may distort projections. In the DDRP
Level II analyses, processes are isolated in order to focus on the principal soil reactions associated
with surface water acidification. Some of the uncertainly in assessing effects is due to this approach
of isolating facets of the whole ecosystem.
9.3.2.4 Model Inputs
The Bloom-Grigal model was designed not to be data intensive. The data required to run the
Bloom-Grigal model fall into four categories: (1) deposition data, (2) precipitation data, (3) soil
chemistry data, and (4) fixed parameters. The sources of the deposition data are described in Section
5. Table 9-7 lists the specific data required to run the Bloom-Grigal model. The soil chemistry data
used in these simulations have been aggregated to a single compartment representing the watershed
204
-------
Table 9-7. List of Input Data for the BIoom-Grigal Soil Acidification Model
Input Variables
Units
Annual average runoff
Annual H+, NH4+, NO3', and SO42" in wet deposition*1
Annual H+, NH4+, NO3", and SO42" in dry deposition3
Soil pHb
Sum of soil exchangeable base cations (0.1 M NH4CI)
Soil cation exchange capacity (0.1 M NH4CI)
cm
keq/ha
keq/ha
PH(H20)
keq/ha
keq/ha
Fixed Parameters
Value
Length of simulation
Partial pressure of soil CO2
Activity coefficient of AI3+
Activity coefficient of AI(OH)2+
100yr
0.005 atm
0.82
0.92
0 Wet and dry SO42" deposition are used to calculate alternative deposition scenarios.
b Determined using deionized water.
205
-------
level. These procedures are described in detail by Johnson et al. (1990). The capacity variables, sum
of exchangeable base cations (SOBC), and CEC are capacity weighted. Soil pH is intensity weighted.
9.3.2.5 Model Outputs
The Bloom-Grigal model simulates soil processes relevant to the assessment of effects of acidic
deposition on soils. During model simulation runs, soil pH, soil base cation status (i.e., base satura-
tion), and soil solution AI3+ are tracked. Each of these is presented in these analyses.
During the 100-year simulations, soil pH, base saturation, and soil solution AI3+ are reported for
years 0, 20, 50, and 100. The projected changes are presented as cumulative distribution functions
(CDF) for graphical comparisons. The CDFs represent regionally weighted projections for soils on the
target population of watersheds. Summary statistics for the CDFs also are presented for numerical
comparisons.
9.3.2.6 Data Sources
The data required to run the Bloom-Grigal model include total annual wet and dry deposition,
total annual runoff, and selected soil chemistry data. These are listed in Table 9-7. All these data
were collected as a part of the DDRP and are discussed in detail in Section 5.
9.3.2.6.1 Deposition data --
The deposition data in this analysis are from two sources. The wet deposition data are
measures of typical year annual loadings. The dry deposition data are estimated, because measures
of dry deposition do not exist for the M-APP. The details on the acquisition/generation of the DDRP
deposition data sets are given in Section 5.6. Because there is some uncertainty associated with
these estimates, particularly in the amount of base cations, we bracketed the dry base cation deposi-
tion data to look at the effect of including 100, 50, or 0 percent of the base cations in base case
estimates of dry deposition. The reductions in dry base cations are offset by concomitant increases in
dry H+ to maintain charge balance. These three deposition datasets are referred to as: (1) Typical
Year (TY), (2) TY Reduced Base Cation (TY-rbc)-TY with 50 percent of the dry base cations, and (3)
TY Zero Base Cation (TY-zbc)-TY with no base cations in dry deposition. A summary of the regionally
weighted median deposition inputs in the three deposition data sets (TY, TY-rbc, and TY-zbc) used in
the Bloom-Grigal modelling in the M-APP is presented in Table 9-8.
Median values, of total acid input (H+tota| = H+ + NH4+ - NO3') in the M-APP are considerably
higher than in the NE and SBRP (compare Table 9-8 with Table 9-24, Church et. al., 1989). Median
acid loadings are at least twice those in the other regions.
9.3.2.6.2 Deposition scenarios -
Three scenarios of future deposition are used in this analysis. They are identical to those
described in Section 9.2.3.2, with the exception of the length of simulation. In lieu of the 140-year
206
-------
Table 9-8. Regionally-Weighted Median Values of Base Year Annual Acid Inputs for the Bloom-
Grlgal Model for Watersheds In the M-APP Region, for Three Levels of Dry Base
Cation Deposition3
TY°
TY-rbc
TY-zbc
H+ NH4+
1.38 0.27
1 .41 0.27
1.55 0.27
NO3" Total Acid lnputb
0.68 0.93
0.68 1 .03
0.68 1.13
a Values are in keq/ha/yr
b Total Acid Input = [H+ + NH4+ - NO3" ]
c TY - typical year wet deposition data for dry deposition; TY-rbc = a 50 percent reduction in loading of base cations in
dryfaN; TY-zfac = no base cations in dry deposition.
207
-------
simulations described in Section 9.2.3.2,100-year simulations are used here. Constant deposition is
referred to as "Scenario A." A timed 50 percent reduction in wet and dry sulfate deposition is referred
to as "Scenario B." A timed 50 percent reduction in wet and dry sulfate deposition followed by a timed
30 percent increase is referred to as "Scenario C." Charge balance in the ramped scenarios is main-
tained by either increasing or decreasing proton deposition.
9.3.2.6.3 Soils data -
The Bloom-Grigal model uses one value for the following soil chemistry variables to depict the
soil chemistry of a particular watershed: soil pH, CEC, and SOBC. To get results that represent the
central tendency of the DDRP regions, a large number of observations for these variables were aggre-
gated to obtain values for each of the DDRP watersheds. Combining or aggregating these data can
be accomplished in several ways. It is probably not correct to use a simple average for all variables;
rather, capacity and intensity variables should be weighted differently. Of the variables used in the
Bloom-Grigal model simulations, soil pH was aggregated using an intensity variable aggregation
method, whereas CEC and SOBC were aggregated using a capacity variable aggregation method.
The details of these methods are provided in Johnson et al. (1990). A summary of the regionally
weighted median values of the Bloom-Grigal soil chemistry input data for the three DDRP regions is
presented in Table 9-9.
Ranking the regional median values for base saturation, the NE is highest followed by the
M-APP and the SBRP. All other things being equal (e.g., soil mass, total acid input, runoff), soil base
saturation provides one means of comparing the proton buffering potential of a soil and the potential
duration of delays in response to acidic deposition. The M-APP, therefore, appears to be intermediate
in buffering potential. We know, however, that this region receives a much larger acid load than the
other two regions. The median soil pH values are low in all three regions, as reflected in the estimates
of AI3+.
9.3.2.7 Results
This section presents results of the Level II base cation analyses using Bloom and Grigal's
(1985) cation depletion model. Following discussions about the input data, the projected effects of
acidic deposition on soil base saturation, soil pH, and soil solution AI3+ are presented. Following
some discussion of these results, comparisons are made to similar projections for the NE and SBRP.
Results for the NE and SBRP are summarized from Church et al. (1989, Section 9.3).
The DDRP was specifically designed to make projections of the regional effects of acidic deposi-
tion. In the DDRP, the basic unit of investigation is the watershed. Instead of characterizing the
effects of acidic deposition on individual soil pedons, the research focus has been the integrated effect
of acidic deposition on watersheds, using a set of randomly selected watersheds that are regionally
representative. As a result, all of the Bloom-Grigal modelling input data and simulation results are at
the watershed level. Because the DDRP sample of watersheds serves as the basic link to the target
population of watersheds, watershed level results are extrapolated to the target population of
watersheds. The statistical design and methods used in the DDRP are described in Section 6.
208
-------
Table 9-9. Regionally-Weighted Median Values of Annual Initial Soil Chemical Values of the
Level II Bloom-Grigal Modelling for the M-APP, NE, and SBRPa
Region
M-APP
NE
SBRP
PH
4.64
4.62
4.85
SOBC
130
40.0
40.4
CEC
958
184
433
BS
13.8
22.0
9.2
AI3+
9.5
10.3
4.3
pH = intensity weighted soil pH; SOBC = mass weighted sum of exchangeable base cations in keq/ha; CEC = mass
weighted cation exchange capacity in keq/ha; BS = base saturation [(SOBC/CEC)*100], in percent; AI3+ = calculated based
upon initial soil conditions in/imol/L
209
-------
9.3.2.7.1 Projections of the effects of acidic deposition on percent base saturation --
The results of the model projections for the effects of acidic deposition on soil base saturation
are presented graphically in Figure 9-13 and in tabular form in fable 9-10 for the three deposition
scenarios. Both Figure 9-13 and Table 9-10 present regionally weighted results for years 0, 20, 50,
and 100 in the simulations. Under current deposition, the regional estimate of percent base saturation
drops from 13.8 percent to 9.4 percent in 100 years. Most of this change is projected to occur within
the first 50 years. This may be indicative of shift in the soil buffer from cation exchange to the
solubilization of aluminum-oxyhydroxides, as suggested by Ulrich (1983). This is considered more
fully in Section 9.3.3.2. -
The effect of the alternate deposition scenarios is clearly seen in Figure 9-13. The middle graph
shows the effect of a ramped 50 percent reduction in sulfate deposition on soil base saturation. The
net effect is seen by the close grouping of the CDF curves. There is little change in soil base satura-
tion after 100 years with deposition scenario B. The change in median base saturation is projected to
be less than 0.7 percent. Deposition scenario C, however, produces a somewhat larger change in
base saturation (the regional median soil base saturation decreased by approximately 2 percent after
100 years), but is less than that projected for constant deposition.
The effects of the two lower concentrations of dry base cations are reported in Tables 9-11 and
9-12. The largest change in base saturation is for the constant deposition scenario. With the TY-rbc
(dry base cations reduced by 50 percent) deposition dataset, base saturations are projected to change
from 13.8 percent to 8.6 percent (Table 9-11), a slightly larger decrease than the final median base
saturation of 9.5 percent for the TY base scenarios. With the TY-zbc (no dry base cation deposition),
the effects are even more dramatic (Table 9-12). The regional median base saturation is decreased by
about half (12.75 percent to 6.7 percent). This represents a worst case, because we know that "there
is some component of base cations in dry deposition. The amount, however, is uncertain.
9.3.2.7.2 Projections of the effects of acidic deposition on soil pH —
Soil pH is a master variable in that it either controls or strongly influences the solubility of soil
minerals, the composition of soil solutions and the cation exchange complex, and soil biochemistry.
Soils are generally well-buffered systems - buffered against changes in pH - due primarily to the
complex nature of soils, which are a mixture of inorganic and organic phases, along with water.
In assessing the effects of acidic deposition on soils, the acid titration of a mixed buffer is a
good conceptual model of the processes involved. As with any buffered system, the buffering
capacity of the system can be exceeded, but because soils are effectively a mixture of several buffers,
as the capacity of one buffer becomes increasingly depleted, the pH of the soil will drop only slightly
as a second buffer takes over. As the acid loading proceeds, the succession of buffer systems
continues. At some point, however, the pH of the system can be reduced to a level at which
dissolution of Al and other metals becomes the primary buffer, and concentrations of these metals
increase to levels that inhibit or are lethal to terrestrial vegetation and aquatic biota.
210
-------
SCENARIO A
1.0
X
VI
o
fc
2
g
CL
0.8-
0.6-
0.4-
0.2"
0.0
10
15 20 25 30
PERCENT BASE SATURATION
35
1.0
SCENARIO B
x -
VI
§
cc
O 0.4-
Q.
g
^ 0.2 H
0.0
10
15 20 25 30
PERCENT BASE SATURATION
35
SCENARIO C
15 20 25 30
PERCENT BASE SATURATION
35
Figure 9-13. Regional CDFs of BIoom-Griga! model projected base saturations of soils on
the DDRP target population of M-APP stream watersheds for three deposition
scenarios (A, B, and C) at 0, 20, 50, and 100 years.
211
-------
Table 9-10. Bloom-Grigal Model Regional Projections of Percent Base Saturation for Soils in the
M-APP Region as a Function of Time and Deposition Scenario9
YEAR
YEAR
YEAR
Deposition Scenario = A
MEAN STD DEV MIN P 25 MEDIAN P 75
Deposition Scenario = B
MEAN STD DEV MIN P 25 MEDIAN P 75
Deposition Scenario = C
MEAN STD DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
16.5
15.5
14.0
12.1
6.0
6.1
6.2
6.5
10.1
9.5
8.2
5.6
12.7
11.9
10.0
8.0
13.8
12.4
10.8
9.4
18.3
18.2
15.4
14.9
34.9
34.4
33.6
32.4
MAX
0
20
50
100
16.5
16.0
15.9
15.8
6.0
6.0
6.0
6.0
10.1
9.9
9.9
9.9
12.7
12.3
12.2
12.2
13.8
13.2
13.2
13.2
18.3
18.3
18.3
18.3
34.9
34.7
34.7
34.7
MAX
0
20
50
100
16.5
16.0
15.6
14.4
6.0
6.0
6.1
6.3
10.1
9.9
9.9
7.7
12.7
12.3
11.8
10.7
13.8
13.2
12.8
11.5
18.3
18.3
18.3
16.3
34.9
34.7
34.6
34.1
Projections made using typical year deposition, including a full complement of base cations in dry deposition, and three
deposition scenarios. Results reported for 20-, 50-, and 100-year projections. The data for year 0 represent initial soil
conditions.
212
-------
Table 9*11. Bloom-Grigal Model Regional Projections of Percent Base Saturation for Soils in
the M-APP Region as a Function of Time and Deposition Scenario and for a
Scenario of a 50 Percent Reduction in Base Cation Dry Deposition8
YEAR
YEAR
YEAR
MEAN
Deposition Scenario = A
STD DEV MIN P 25 MEDIAN P 75
Deposition Scenario = B
MEAN STD -DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
16.5
15.2
13.3
11.1
6.0
6.0
6.1
6.2
10.1
9.3
7.8
5.2
12.7
11.7
9.7
7.4
13.8
12.3
10.4
8.6
18.3
17.5
14.3
13.4
34.9
34.1
33.0
31.1
MAX
0
20
50
100
16.5
15.8
15.6
15.3
6.0
6.0
6.0
6.0
10.1
9.8
9.8
9.4
12.7
12.2
12.0
11.6
13.8
13.0
12.9
12.4
18.3
18.0
18.0
17.7
34.9
34.5
34.5
34.4
Deposition Scenario = C
MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
16.5
15.8
15.2
13.5
6.0
6.0
6.0
6.1
10.1
9.8
9.7
7.1
12.7
12.2
11.3
9.5
13.8
- 13.0
12.1
10.7
18.3
18.0
17.5
14.9
34.9
34.5
34.1
32.9
Projections made using typical year deposition, with three scenarios of future deposition. Results reported
for 20-, 50-, and 100-year projections. The data for year 0 represent initial soil conditions.
213
-------
Table 9-12. Bloom-Grigal Model Regional Projections of Percent Base Saturation for Soils in the
M-APP Region as a Function of Time and Deposition and for a Scenario of No Base
Cations in Dry Deposition8
YEAR
YEAR
YEAR
MEAN
Deposition Scenario = A
STD DEV MIN P 25 MEDIAN P_75
MAX
0
20
50
100
16.5
14.9
12.7
10.1
6.0
6.0
6.0
5.9
10.1
9.2
7.5
4.8
12.7
11.4
8.8
6.8
13.8
12.1
10.0
7.7
18.3
16.8
13.3
12.1
34.9
33.9
32.3
29.8
Deposition Scenario = B
MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
16.5
15.6
15.2
14.7
6.0
6.0
5.9
5.8
10.1
9.7
9.7
8.8
12.7
12.1
11.7
11.0
13.8
12.7
12.2
12.2
18.3
17.7
17.7
16.2
34.9
34.2
33.8
33.1
MEAN
Deposition Scenario = C
STD DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
16.5
15.6
14.2
12.5
6.0
6.0
5.9
5.9
10.1
9.7
9.3
6.6
12.7
12.1
10.9
8.8
13.8
12.7
11.6
9.9
18.3
17.7
16.4 .
14.2
34.9
34.2
33.4
31.6
Projections made using typical year deposition, with three scenarios of future deposition. Results reported for 20-, 50-, and
100-year projections. The data for year 0 represent initial soil conditions.
214
-------
The Bloorn-Grigal modelling results of the effects of acidic deposition on soil solution pH in the
M-APP uplands are summarized in Figure 9-14 and Tables 9-13, 9-14, and 9-15. The soil pH levels in
the M-APP are initially low from an agricultural point of view, but are not, however, exceptionally low
for forested soils. The aggregated soil pH values for the target population of watershed systems in the
M-APP ranged from about 4.5 to 5.3 (Table 9-13). In developing the aggregate soil pH values, the
organic horizons were included. Because organic horizons are often the most acidic horizons in
forested soils, their contribution to the aggregate pH is reflected in these low regional values.
Under constant deposition (Scenario A) the Bloom-Grigal model projects that the median soil pH
in the M-APP will decrease by 0.2 pH units over the next 100 years (Table 9-13). The systems with a
current pH above 4.7 (approximately 10 percent of the systems in the target population; Figure 9-14)
appear to be strongly buffered against change, and change only slightly if at all. Approximately 90
percent of the systems (those with initial aggregate soil pH values less than 4.7) are expected to show
decreases in soil pH under constant deposition at current levels. Although the projected changes in
pH are not large, they represent significant increases in H+ concentration in the soil solution, and
result in significant increases in soil solution aluminum. The relatively small changes in pH, compared
to those for exchangeable base cations and soil solution aluminum, in part reflect the reporting of a
logarithmic variable, but they result mainly from (1) selectivity coefficients that favor exchange of H+
from liquid to solid phase (with release of base cation from the exchange complex) and (2) buffering of
pH changes by dissolution of aluminum. The changes in pH and Al could have significant adverse
consequences, both for terrestrial vegetation and for biota of receiving surface waters.
The effect on model projections for pH of using the alternate dry base cation deposition scen-
arios is small. For deposition scenario A (constant deposition at current loadings), use of the TY-rbc
(Table 9-14) and TY-zbc (Table 9-15) deposition data results in only slightly lower projected soil pH
values than straight TY data (Table 9-13). Decreasing the base cations in dry deposition results in
increased acid loadings to the soil because H+ is increased to maintain change balance. As a result,
the rate of change in base saturation is slightly faster than in the base case, but there is almost no
effect of soil pH, because pH is buffered by the soil (Equation 9-6). Thus the Bloom-Grigal model
projections in almost all cases include much larger changes in base cations than in pH, and model
projections are relatively insensitive to small changes (such as changes in dry base cation deposition)
in base cation flux. Figure 9-15 shows the relationship between-soil pH and base saturation for typical
values of n and pKa. The flat part of the curve represents conditions under which the soil is buffered
against changes in soil pH, such that even large changes in base saturation are associated with minor
changes in soil pH. The steeper portions of the curve, where the slope is greatest, are the condition
of the soil titration curve, where small changes in soil base saturation result in large changes in soil
pH. This represents a shift from one soil pH buffering mechanism to another, such as a shift from a
cation exchange buffer to an aluminum buffer (Ulrich, 1983).
As with the soil base saturation projections reported in the foregoing discussion, the model-
based projections of change in soil pH under decreased deposition (Scenario B) result in only slight
changes in soil pH (Table 9-13 and middle graph on Figure 9-14). Even though the projected pH
change after 100 years under Scenario B is small, most of it occurs in the first 20 years of the
215
-------
SCENARIO A
X
VI
~z.
o
cc
o
CL
O
cc
CL
1.0-
0.8-
0.6-
0.4-
0.2-
o.o -
4.
I
1
^
*
*
•
t
1
1
!
,' /
i i
i / /
' i
I Jf*
'" 1
* — ' /
i i
DO 4.25 4.5(
..^^^^^^f^ff^'
• ft*
/ 1 '"•
1 .'
1 .
1 •
i;
i .'
/ •
i t
•
/ * ----- 0 Year
i* 20 Years
50 Year"
. "jQQ Years
i i i
) 4.75 5.00 5.25 5.!
SOIL pH
SCENARIO B
X
VI
z
o
1-
cc
o
CL
o
cc
CL
i.o-
0.8-
0.6-
0.4-
0.2-
0.0"
4.
/T"
!•'
y •
H "
r
yf
|r
/*
V*
f f
/
J
/
/•
DO 4.25 4.50 4.75 5
.^— -ir^*^
20 Years
50 Year"
1 00 Years
00 5.25 5.
50
SOIL pH
SCENARIO C
X
VI
•z.
O
o
0-
o
cc
Q.
1.0-
0.8-
0.6-
0.4-
0.2"
0.0 "
4.
,.^£5***""
jjT
/ // »
/ /i *
/ /i."
;' '•*
|/n
(1 1 OYear
IV
> 20 Years
' / 50 Year*"
'r 1Afl Yea re
/^ — — — luU I tJafo
,-
DO 4.25 4.50 4.75 5.00 5.25 5.!
SOIL pH
Figure 9-14. Regional CDFs of Bloom-Grigal model projected pH of soils on the DDRP
target population of M-APP stream watersheds for three deposition scenarios
(A, B, and C) at 0, 20, 50, and 100 years, using deposition data for the full
complement of base cations in dry deposition.
216
-------
Table 9-13. Bloom-Grigal Model Regional Projections of Soil Solution pH for Soils in the M-APP
Region as a Function of Time and Deposition Scenario for a Scenario of a Full
Complement of Dry Base cations in Dry Deposition8
YEAR
YEAR
MEAN
Deposition Scenario = A
STD DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
4.7
4.6
4.6
4.5
0.2
0.2
0.2
0.2
4.5
4.5
4.4
4.2
4.6
4.7
4.5
4.4
4.6
4.6
4.5
4.4
4.7
4.6
4.6
4.5
5.3
5.3
5.3
5.2
Deposition Scenario = B
YEAR MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
4.7
4.6
4.6
4.6
0.2
0.2
0.2
0.2
4.5"
4.5
4.5
4.4
4.6
4.6
4.6
4.6
4.7
4.6
4.6
4.5
4.7
4.7
4.6
4.6
5.3
5.3
5.3
5.3
Deposition Scenario = C
MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
4.7
4.6
4.6
4.6
0.2
0.2
0.2
0.2
4.5
4.5
4.5
4.4
4.6
4.6
4.6
4.5
4.6
- 4.6
4.6
4.5
4.7
4.7
4.6
4.6
5.3
5.3
5.3
5.3
0 Projections made using typical year deposition, with three deposition scenarios. Results reported for 20-,
50% and 100-year projections. The data for year 0 represent initial soil conditions.
217
-------
Table 9-14. Bloom-Grigal Model Regional Projections of Soil Solution pH for Soils in the M-APP
Region as a Function of Time and Deposition Scenario and for a Scenario of a 50
Percent Reduction in Base Cations in Dry Deposition8
YEAR
YEAR
YEAR
MEAN
Deposition Scenario = A
STD DEV MIN P 25 MEDIAN P 75
MEAN
Deposition Scenario = C
STD DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
4.7
4.6
4.6
4.5
0.2
0.2
0.2
0.2
4.5
4.5
4.4
4.2
4.6
4.6
4.5
4.4
4.6
4.6
4.5
4.4
4.7
4.6
4.5
4.5
5.3
5.3
5.2
5.2
Deposition Scenario = B
MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
4.5
4.6
4.6
4.6
0.2
0.2
0.2
0.2
4.5
4.5
4.5
4.5
4.6
4.6
4.6
4.5
4.6
4.6
4.6
4.6
4.7
4.7
4.6
4.6
5.3
5.3
5.3
5.3
MAX
0
20
50
100
4.7
4.6
4.6
4.6
0.2
0.2
0.2
0.2
4.5
4.5
4.5
4.4
4.6
4.6
4.5
4.5
4.6
4.6
4.6
4.5
4.7
4.7
4.6
4.6
5.3
5.3
5.3
5.2
a Projections made using typical year deposition, with three deposition scenarios. Results reported for 20-,
50-, and 100-year projections. The data for year 0 represent initial soil conditions.
218
-------
Table 9-15. BIoom-Grigal Model Regional Projections of Soil Solution pH for Soils in the M-APP
Region as a Function of Time and Deposition Scenario and for a Scenario of Zero
Base Cations in Dry Deposition3
YEAR
YEAR
MEAN
Deposition Scenario = A
STD DEV MIN P 25 MEDIAN P 75
MEAN
Deposition Scenario = C
STD DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
4.7
4.6
4.5
4.4
0.2
0.2
0.2
0.2
4.5
4.4
4.4
4.2
4.6
4.5
4.4
4.3
4.6
4.6
4.5
4.4
4.7
4.6
4.5
4.4
5.3
5.3
5.2
5.2
Deposition Scenario = B
YEAR MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
4.7
4.6
4.6
4.6
0.2
0.2
0.2
0.2
4.5
4.5
4.5
4.4
4.6
4.6
4.5
4.5
4.6-
4.6
4.6
4.6
4.7
4.7
4.6
4.6
5.3
5.3
5.3
5.2
MAX
0
20
50
100
4.7
4.6
4.6
4.5
0.2
0.2
0.2
0.2
4.5
4.5
4.5
4.3
4.6
4.6
4.5
4.4
4.6
4.6
4.6
4.5
4.7
4.7
4.6
4.5
5.3
5.3
5.3
5.2
Projections made using typical year deposition, with three deposition scenarios. Results reported for
20-. 50-, and 100-year projections. The data for year 0 represent initial soil conditions.
219
-------
8
2-
0
0.0
0.2 0.4 0.6 0.8
SOIL BASE SATURATION
1.0
pH = pK + n log [BS/(1-BS)]
pKa = 4.96 n = 0.797
Figure 9-15. The curvilinear relationship between soil pH and base saturation for typical
values of n and pKa as defined by Equation 9-6.
220
-------
simulations. In all likelihood, this occurs because of the nature of the deposition scenario. For the
first five years of the simulation, the acid loadings are constant and equal to current levels. Therefore,
the projected trajectory soil pH is the same for deposition scenarios A and B for the first five years.
For the next 10 years under scenario B, however, the deposition is ramped down by 5 percent per
year, until the initial deposition is halved, then held constant for the duration of the simulation. Under
scenario B, the rate of decrease in soil pH slows beginning in year 6 and is essentially flat at year 15
and for the remainder of the simulation.
As expected, the projected changes for soil pH under deposition scenario C (ramped down to
50 percent of current loadings but ramped back up to 80 percent) are less than those for scenario A
but greater than under scenario B. The conclusion to be drawn from this analysis is that reducing
acid loadings diminishes the projected effects of acidic deposition on soil pH. A corollary can be
drawn for the concomitant effects on vegetation and surface waters.
9.3.2.7.3 Projections of the effects of acidic deposition on estimates of soil solution AI3+ --
It has been known for some time in agricultural systems that low soil pH can result in increased
soil solution Al concentrations that are potentially toxic to plants. Cronan and Schofield (1979)
proposed that Al could be mobilized by acidic deposition and subsequently transported to surface
waters where aquatic organisms could be affected. Increasing soil solution Al concentrations as a
result of acidic deposition has also been suggested as a factor that may contribute to forest stress and
decline (Ulrich, 1983). Although Al toxicity levels are not known for very many forest species, we
assume that increases in soil solution Al will have adverse effects upon forest vegetation, and will lead
to increased transport of Al to surface waters. The purpose of this analysis is use the Bloom-Grigal
model projections of soil solution AI3+ concentrations to evaluate the effects acidic deposition on soil
solution Al3* in the M-APP.
The results of this analysis are presented in Figure 9-16, and in Tables 9-16, 9-17, and 9-18.
Under deposition scenario A (Table 9-16), median soil solution AI3+ concentrations are projected to
more than double during the 100-year simulation. The initial median concentration is 9.5^mol/L; it
increases to 22 .umol/L at the end of the simulation. Approximately 10 percent of the systems are
projected to have AI3+ concentrations in excess of 30 ^mol/L at the end of 200 years. Reductions in
deposition, represented by scenario B, have a dramatic effect, reducing the projected concentration of
Al3* in solution at 100 years by half. In deposition scenario C, the model projects less AI3+ in
solution compared to scenario A, but slightly more than scenario B. In all cases, decreasing the
estimated concentration of base cations in dry deposition resulted in estimates of lower pH and
greater projected AI3+ concentrations relative to otherwise identical scenarios.
It is apparent from this analysis that the soils in the M-APP are likely to have increased AI3+
concentrations at current levels of deposition. This situation increases the likelihood that surface
waters will have increased AI3+ concentrations that may adversely affect aquatic biota.
221
-------
1.0
X
VI
o
0.8-
0.6-
a:
O 0.4
CL
O
cc
°- 0.2
0.0
SCENARIO A
10
1 1
20 30
AI3+(nmol/L)
40
50
1.0
0.8-
X
VI
•Z. 0.6-
O
t-
CC
O 0.4-
CL
g
CL 0.2-
0.0
SCENARIO B
10
20 30
A!3+(iamol/L)
40
50
1.0
0.8-
:o.6-
CC
00.4
O
cc
0.0
SCENARIO C
10
—r
20
30
AI3+(Hmol/L)
40
50
Figure 9-16. Regional CDFs of Bloom-Grigal model projected soil solution concentrations
of AI3+ of soils on the DDRP target population of M-APP stream watersheds
for three deposition scenarios (A, B, and C) at 0, 20, 50, and 100 years.
222
-------
Table 9-16. EJIoom-Grigal Model Regional Projections of Soil Solution AI3+ for Soils in the
M-APP Region as a Function of Time and Deposition Scenario for the Scenario of
a>. Full Complement of Base Cations in Dry Deposition3
YEAR
YEAR
YEAR
MEAN
Deposition Scenario = A
STD_DEV MIN P 25 MEDIAN P 75
MEAN
Deposition Scenario = B
STD DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
10.3
11.7
14.3
20.0
4.0
4.2
4.9
7.8
0.8
0.9
0.9
1.0
8.0
9.8
12.8
15.0
9.5
11.4
14.6
22.0
11.9
13.0
16.3
25.6
17.3
19.2
23.3
44.5
MAX
0
20
50
100
10.3
10.9
11.0
11.2
4.0
4.1
4.0
4.0
0.8
0.9
0.9
0.9
8.0
8.9
9.6
9.6
9.5
10.3
10.3
10.7
11.9
12.2
12.2
12.6
17.3
18.0
18.0
18.0
Deposition Scenario = C
MEAN STD_DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
10.3
10.9
11.5
13.6
4.0
4.1
4.0
4.5
0.8
0.9
0.9
0.9
8.0
9.0
9.9
11.0
9.5
10.3
11.1
14.8
11.9
12.2
13.5
16.1
17.3
18.0
18.0
25.9
Projections made using typical year deposition, with three deposition scenarios. Results reported for 20-,
50-, and 100-year projections. The data for year 0 represent initial soil conditions. Units are /jmol/L.
223
-------
Table 9-17. BIoom-Grigal Model Regional Projections of Soil Solution AI3+ for Soils in the
M-APP Region as a Function of Time and Deposition Scenario for the Scenario of
a 50 Percent Reduction of Base Cations Concentration in Dry Deposition3
YEAR
YEAR
YEAR
MEAN
Deposition Scenario = A
STD DEV MIN P 25 MEDIAN P 75
MEAN
Deposition Scenario = B
STD DEV MIN P 25 MEDIAN P 75
MEAN
Deposition Scenario = C
STD DEV MIN P 25 MEDIAN P 75
MAX
0
20
50
100
10.3
12.1
15.7
23.7
4.0
4.4
5.4
9.5
0.8
0.9
1.0
1.1
8.0
10.1
14.1
18.1
9.5
11.4
15.4
24.S
11.9
13.3
18.8
28.9
17.3
19.7
25.1
50.2
MAX
0
20
50
100
10.3
11.2
11.4
11.9
4.0
4.1
4.1
4.1
0.8
0.9
0.9
0.9
8.0
9.1
9.8
10.2
9.5
10.6
10.6
11.3
11.9
12.4
12.7
14.1
17.3
18.2
18.2
18.4
MAX
0
20
50
100
10.3
11.5
12.1
15.3
4.0
4.1
4.1
5.2
0.8
0.9
0.9
1.0
8.0
9.1
10.6
12.6
9.5
10.6
11.6
16.1
11.9
12.4
14.2
19.5
17.3
18.2
18.5
29.4
Projections made using typical year deposition, with three deposition scenarios. Results reported for 20-
50-, and 100-year projections. The data for year 0 represent initial soil conditions.
224
-------
Table 9-18. BIoom-Grigal Model Regional Projections of Soil Solution AI3+ for Soils in the
M-APP Region as a Function of Time and Deposition Scenario for the Scenario of
Zero Base Cations in Dry Deposition9
Deposition Scenario = A
YEAR MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
10.3
12.5
17.3
28.2
4.0
4.7
6.1
11.8
0.8
0.9
1.0
1.2
8.0
10.4
14.6
22.0
9.5
12.4
16.8
29.2
11.9
13.7
21.9
34.1
17.3
20.3
26.9
56.4
Deposition Scenario = B
YEAR MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
10.3
11.5
11.9
12.8
4.0
4.3
4.2
4.4
0.8
0.9
0.9
1.0
8.0
9.3
10.2
10.7
9.5
11.0
11.2
12.4
11.9
12.8
14.1
16.7
17.3
18.5
18.5
20.5
YEAR
Deposition Scenario = C
MEAN STD DEV MIN P 25 MEDIAN P 75 MAX
0
20
50
100
10.3
11.5
12.9
17.8
4.0
4.3
4.4
6.3
0.8
0.9
0.9
1.1
8.0
9.3
11.2
15.0
9.5
11.0
12.2
18.1
11.9
12.8
15.8
21.8
17.3
18.5
19.3
33.3
Projections made using typical year deposition, with three deposition scenarios. Results reported for 20-,
50-, and 100-year projections. The data for year 0 represent initial soil conditions.
225
-------
9.3.2.7.4 Summary --
Characteristics of soils in the M-APP uplands and model projections of future chemical changes
in soils, using the Bloom-Grigal soil acidification model, can be summarized as follows:
• Soils in the M-APP do not presently have extremely low base saturations or phi, or extremely
high estimated AI3+ concentrations in comparison to other forested temperate soils.
• Using current deposition in the Bloom-Grigal model to make projections for a 100-year
period, the regional median soil base saturations are projected to drop below 10 percent
(13.8 percent in year 0, 12,4 percent, 10.8 percent and 9.4 percent at years 20, 50, and
100).
• Under a scenario of constant deposition, regional median soil solution AI3+ concentrations
are projected to more than double, from 9.5/
-------
Bloom-Grigal model projections were made using the respective regional soils and deposition
data. The deposition data used in the NE and SBRP were long-term annual average estimates of
current wet and dry deposition (Church et al., 1989, Section 5.6), whereas the M-APP projections are
based on a typical year data set as described in Section 5.6 and Section 9.3.2.1. The median annual
total acid inputs for the NE, M-APP, and SBRP were 0.43, 0.93, and 0.47 Keq/ha, respectively. With
these estimates of current deposition and the regional soils data, the Bloom-Grigal model was used to
project potential! changes in soil and soil-solution chemistry. The results were regionally weighted and
summarized quinquennially over a 100-year simulation period. The median values projected for base
saturation, soil pH, and soil solution AI3+ are plotted for each region as a function of time in Figures
9-17, 9-18, and 9-19, respectively, and are summarized in Table 9-19.
9.3.2.8.1 Projections of the effects of acidic deposition on soil base saturation -
The regional median value of base saturation initially is highest for the NE followed by the
M-APP and then the SBRP (Table 9-19). This north-to-south ranking is expected based upon the age
and degree of v/eathering of the soils in each of the regions. The NE is the youngest landscape of the
three, with soils roughly 15,000 years old. The soils in the NE tend to be Entisols, Inceptisols, and
Spodosols. The soils in the M-APP are older; most are Inceptisols, Entisols, or Alfisols. The soils in
the SBRP tend to be the highly weathered Ultisols.
Figure 9-17 shows the projected median base saturations as a function of time. Median base
saturation in soils in each of the three regions decreases somewhat after 100 years; the greatest
projected change occurs in M-APP soils. This is due largely to the higher loading of acidic deposition
in the M-APP. The NE appears to be relatively well buffered against changes in soil base saturation.
Soils in the SBRP start at approximately 9 percent base saturation and decrease to slightly less than 7
percent at the end of the simulation. The final projected base saturations in the M-APP and SBRP
soils are low enough that the fraction of AI3+ in soil leachate can increase significantly, and macro-
nutrients may become limiting to terrestrial vegetation.
9.3.2.8.2 Projections of the effects of acidic deposition on soil pH -
The median pH values of soils in the NE and M-APP initially are similar and slightly lower than in
SBRP soils (Table 9-19). After 100 years, simulated median pH in the M-APP and SBRP soils
decreases by about 0.2 pH units. The changes in pH are relatively small, but represent large
increases in H+, and are projected to be accompanied by substantial increases in soil solution AI3+.
Soils in the NE Region, due to their relatively high base status, are more strongly buffered, and
experience smaller decreases in soil pH. Figure 9-18 graphs these results.
9.3.2.8.3 Projections of the effects of acidic deposition on estimates of soil solution AI3+ --
Soil solution AI3+ results were not presented in the earlier DDRP report on the NE and SBRP
(Church et al., 1989). The Bloom-Grigal model was rerun on these two regions using the methods,
data, and protocols described by Church et al. (1989). The results of these runs, along with those for
the M-APP are presented here to allow regional comparisons.
227
-------
25
O
< 20 -j
DC
CO
LLJ
CO
CD
r-
"Z.
LLI
O
or
LJJ
Q_
15-
10-
0
YEAR 0 = 1985 YEAR 50 = 2035 YEAR 100 = 2085
NE
0
i
10
20 30 40 50 60 70 80
YEARS IN SIMULATION
90 100
Figure 9-17. Projected changes in median base saturation for 100-year simulations, for
soils in DDRP target population watersheds in the NE, M-APP, and SBRP
regions, generated using the Bloom-Grigal model for a scenario of constant
deposition at current levels.
228
-------
5.0
4.8-
31
Q.
O
CO
4.6-
4.4-
4.2
YEAR 0 = 1985 YEAR 50 = 2035 YEAR 100 = 2085
0 10 20 30 40 50 60 70 80 90 100
YEARS IN SIMULATION
Figure 9-18. Projected changes in median pH, for 100-year simulations, for soils in DDRP
target population watersheds in the NE, M-APP, and SBRP regions, generated
using the BIoom-Grigal model for a scenario of constant deposition at current
levels.
229
-------
25
YEAR 0 = 1985 YEAR 50 = 2035 YEAR 100 = 2085
30 40 50 60 70
YEARS IN SIMULATION
90 100
,3+
Figure 9-19. Projected changes in median soil solution AT, for 100-year simulations, for
soils in DDRP target populations in the NE, M-APP, and SBRP regions, gener-
ated using the Bloom-Grigal model for scenarios of constant deposition at
current levels.
230
-------
Table 9-19. Summary Statistics for Regional Medians for Base Saturation (BS), pH, and Soil
Solution AI3+ Concentration for Soils on DDRP Target Populations of Watersheds
In the NE, M-APP, and SBRPa
CONSTANT DEPOSITION
Region
YEAR IN SIMULATION
Parameter
20
50
100
NE
M-APP
SBRP
%BS
%BS
%BS
22.0
13.8
9.2
21.7
12.4
8.1
21.2
10.8
7.3
21.0
9.4
6.6
NE
M-APP
SBRP
PH
pH
PH
4.6
4.6
4.9
4.6
4.6
4.8
4.6
4.5
4.7
4.6
4.4
4.7
NE
M-APP
SBRP
Ald
Al1
3+
10.3C
9.5b
4.3b
10.7
11.4
5.2
11.2
14.6
6.3
12.2
22.0
7.9
0 Projections were made using the Bloom-Grigal model, for scenarios of constant future deposition.
b ;
-------
Soil solutions in the NE have the highest projected initial soil solution AI3+ concentrations (Table
9-19). After 100 years, however, the median AI3+ levels are projected to be greatest in the M-APP. Al
concentrations increase slightly in NE soil solutions and in the SBRP they increase somewhat more
than in the NE, as shown in Figure 9-19. There may be several reasons for these results. First, the
acid inputs in the M-APP are approximately twice those in the other two regions, resulting in more
extensive cation leaching from soils and thus larger absolute decreases in base saturation and soil pH
in the region. The second apparent factor is differences in selectivity of exchange reactions for the
two regions, probably resulting from the nature and amount of organic and clay materials in soils of
the three regions. The soils in the NE have low pH, but relatively high base saturations. The SBRP
soils have low base saturations, but the highest pH. The soils in the M-APP are low both in pH and
exchangeable base cations. These characteristics in combination lead to higher projected levels of
Al3+, and consequently indicate that the surface waters in the M-APP are likely to be more susceptible
to acidification than those in the NE and SBRP.
9.3.2.8.4 Summary and conclusions -
Using the Bloom-Grigal model to make projections of the effects of acidic deposition on soils in
the NE, M-APP, and SBRP, and using the results to make regional comparisons, the following
conclusions can be drawn:
• Soils in the NE appear to be buffered against large changes in soil pH, base saturation, and
soil solution AI3+ concentrations under current levels of deposition over the next 100 years.
• The M-APP appears to be susceptible to adverse changes in soil properties due to acidic
deposition at current levels over the next 100 years. Soil base saturation and soil pH are
both projected to decrease. At the same time, soil solution AI3+ concentrations are
projected to more than double.
• The soils in the SBRP are projected to have decreases in base saturation and soil pH
comparable to those in the M-APP. Projected changes in soil solution AI3+ concentrations
are substantially smaller than those in the M-APP.
Based upon model projections, the M-APP appears to be the most susceptible of the three regions, to
detrimental changes in soil properties due to acidic deposition.
9.3.3 Reuss Model
9.3.3.1 Model Description
The Reuss model was originally developed by Reuss (1983) and coworkers (Reuss and
Johnson, 1985; Johnson and Reuss, 1984) to simulate changes in soil and soil solution chemistry
associated with changes in mobile anion leaching, pCO2, etc. In this equilibrium-based, mass balance
model, the solubility of a gibbsite-like phase is assumed to control the concentration of aluminum.
Concurrently, exchange reactions are used to partition the cations AI3+, Ca2+, Mg2+, Na+, and K+
between the solid and solution phases. Figure 9-20 presents schematically the processes considered
232
-------
Top
Horizon
n
Middle
Horizon(s)
Bottom
Horizon
Figure 9-20. Row chart showing hydrologic pathways and soil geochemical processes con-
sidered by the modified Reuss model used for DDRP simulations. Arrows
indicate major pathways through which ions are interchanged among reser-
voirs. No attempt was made to distinguish relative fluxes among reservoirs;
the heavier lines indicate those processes that are the focus of Level II
modelling efforts discussed here. SEC represents soil exchangeable base
cations.
233
-------
in the model. The model computes soil pH, soil solution ANC, and base cation and dissolved
aluminum concentrations. The model then "re-equilibrates" soil solutions with atmospheric carbon
dioxide and computes surface water composition. The development and process representation of the
model, as well as the rationale for its use in the DDRP and a discussion of modifications of the model
for use in the DDRP, are described in detail by Church et al. (1989, Section 9.3).
Reuss's approach has several advantages over the use of simple exchange reactions for model-
ling exchange reactions in soil environments. First, the charge balance requirement of the code
makes the model responsive to ionic strength. Composition of the soil solutions in forested soils
has been shown to depend on ionic strength (Richter et al., 1988). Therefore, this aspect of the model
permits a more realistic simulation of natural exchange reactions than do the less involved computa-
tions. Second, the model allows the user to specify the partial pressure of carbon dioxide (pCO2) in
the soil gas. Although the annual maximum pCO2 in forested soils is typically about 1 percent
(Fernandez and Kosian, 1987; Solomon and Cerling, 1987), this concentration is about 30 times higher
than atmospheric pCO2 and can significantly affect soil solution composition (Reuss and Johnson,
1985, 1986). Third, although use of a gibbsite-like phase to regulate aluminum activities is contro-
versial (c.f., Section 9.3.2.2.7), its use allows one degree of freedom in the solution composition
to be effectively constrained. Finally, the mass balance constraints allow the user to track cation
depletion from the exchange complex as a function of time, hydrogen ion loading, and the imposed
physico-chemical environment.
The Reuss model focuses on soil exchange reactions. The model does not consider other
cation source/sink processes such as mineral weathering, nitrogen transformations, or afforestation,
even though these processes may have equal or greater influence in regulating surface water compo-
sition in certain ecological settings. Other such cation sources/sinks could be incorporated into the
model, but have been excluded from the present analyses so that the potential role of cation exchange
in the soil as a long-term buffer of acidic deposition can be assessed more effectively. Models have
been developed that include these processes, and provide integrated projections of system response
to the imposed deposition (Cosby et al., 1985a,c, 1986a,b; Gherini et al., 1985; Galloway et al., 1983;
Nikolaidis et al., 1988). These integrated models provide a more complete assessment of ecosystem
response, but cannot be used effectively to understand the contributions of individual processes, such
as soil cation exchange, to the buffering responses of watersheds.
This section describes Level II analyses of cation exchange buffering for DDRP watersheds in
the Mid-Appalachian Region, conducted using the Reuss model. As in comparable analyses for DDRP
systems in the NE and SBRP, our intent is to provide a quantitative assessment of the time frame over
which cation exchange processes can neutralize acidic deposition, at current or altered rates of
deposition. Consistent with the objectives of the Level II analyses, these analyses were designed to
consider only exchange processes, and only those occurring in the top two meters of the regolith. No
mineral weathering or other inputs of base cations are considered, except for atmospheric deposition.
As such, these analyses represent a worst-case analysis of exchange-mediated buffering in the soil,
designed to place an upper bound on the potential for the changes in soil and runoff chemistry that
result from exchange-mediated neutralization of acidic inputs by exchange processes within the soils
of DDRP watersheds.
234
-------
9.3.3.2 Model Formulation
Details of the modification to, and resulting evolution of, Reuss's model as used in this study,
have been described in detail by Church et al., 1989. The following primary changes were incor-
porated into the model from the original formulation of Reuss (1983): (1) the model is coded in
FORTRAN instead of BASIC, (2) the model employs the Vanselow exchange formulation (Vanselow,
1932) instead of the Gaines-Thomas formulation (Gaines and Thomas, 1953), (3) nitrogen species are,
to a large extent, presumed to be retained by the ecosystem and are not considered as mobile
species, (4) slightly different aluminum speciation algorithms are used, (5) multiple horizons, which
permit lateral runoff, are incorporated into the model structure, and (6) the DDRP version of the model
employs time steps in the computations instead of water increment steps. The decision to substitute
the Vanselow exchange expression for the Gaines-Thomas expression used in the original Reuss
model formulation was based on two criteria. First, although possible problems remain with the
definition of solid phase activity coefficients, the Vanselow expression is believed by many researchers
to be the most thermodynamically correct exchange formulation. Second, and more important, in
model simulations conducted as part of DDRP analyses for the NE/SBRP to compare model results
using Gaines-Thomas, Vanselow, and Gapon expressions, there were not strong differences among
the three formulations. Residuals generated using the Vanselow formulation were better behaved than
those from the other models.
9.3.3.3 Assumptions
As with any model, assumptions are necessary regarding certain processes, the soil environ-
ment, and the characteristics of certain reactions. The reader is referred to the NE and SBRP DDRP
report (Church, et al., 1989) for a detailed assessment of the effect these assumptions have on model
predictions. The primary assumptions for current analyses include:
• Gibbsite solubility controls soil AI3+ concentrations. This is different from the assumption of
Al control by exchange equilibrium in the Bloom-Grigal model used in Section 9.3.2, and is
a source of potential differences in model simulation results;
• Exchange selectivity coefficients are assumed to be constant, i.e., not to vary with changes
in base saturation; and
• Soil gas pCO2 is assumed to be a uniform 0.005 atm for all soils for all seasons.
9.3.3.4 Limitations
The Reuss model focuses on soil exchange reactions. The model does not consider other
processes such as sulfate adsorption, mineral weathering, nitrogen transformations, or afforestation,
even though these processes may have equal or greater influence in regulating surface water compo-
sition in certain ecological settings (Likens et al., 1977; Johnson et al., 1988). For the present
application of assessing the potential for delays in acidification attributable to soil cation exchange
processes, these limitations are not only appropriate, but necessary. Some of the most significant
omissions in the model are: (1) no provisions are made to consider organo-cations, and especially
235
-------
organo-aluminum interactions, (2) mineral weathering is not considered, with the result that the pool of
base cations in the soil and the relative abundance of individual cation is controlled solely by
deposition and exchange equilibria, and (3) reactions are modelled as equilibrium processes. Clearly,
most ongoing chemical processes in watersheds are subjected to rapidly and constantly changing
chemical environments, but given the rapid nature of exchange reactions and the annual time step
used in simulations, an assumption of exchange equilibrium is not unreasonable.
9.3.3.5 Model Inputs
9.3.3.5.1 Deposition and associated data ~
The model requires deposition data, including precipitation quantity (cm/yr) and average annual
concentrations of the major ions (SO42", CI", NO3', NH4+, Ca2+, Mg2+, Na+, and K+) in precipitation,
along with estimated fluxes of the same ions in dry deposition. The atmospheric flux of each ion is the
combined wet plus dry average annual deposition. Evapotranspiration (% ET) data required by the
model are computed as the quantity of precipitation minus runoff. Sources of deposition, precipitation,
and runoff data are described in Section 5.6 and 5.7.
9.3.3.5.2 Soils data -
The model requires physical and chemical information about each of the horizons included in
the simulations. Physical parameter data used in the model include horizon thickness, bulk density,
percent coarse fragments, and a hydrologic runoff parameter. Chemical parameters include CEC,
base cation concentrations on the exchange complex, selectivity coefficients for the Ca/AI, Ca/Mg,
Ca/Na, and Ca/K exchange reactions, soil gas pCO2, the apparent solubility product for AI(OH)3(s),
and the stoichiometric coefficient for H+ to be used in describing the dissolution of the aluminum solid
phase. Multiple horizon versions of the model require the above information for each of the horizons
to be considered. Generation of soil physical and chemical data are described in Section 5.4 and 5.5.
Church et al. (1989, Section 9.3) provide a detailed description of the procedures used in calculating
exchange selectivity coefficients; procedures used for data aggregation are described by Johnson et
al. (1990).
9.3.3.6 Results - ANC Projections
For each simulation, the model generates projections for surface water composition and, on a
by-horizon basis, for soil and soil solution composition for major chemical species. Results are com-
piled for the first and final years of the computation and at user-specified intervals during the
simulation.
Information on surface water variables retained in the output files include data on pH, ANC,
SO42', NO3",Cr, Ca2+, Mg2+, Na+, K+, AI3+, sum-(AI)aq, and ionic strength. For soils, information on
soil pH, base saturation, and exchangeable Ca, Mg, Na, and K are retained for the solid phase, and
ANC, Ca2+, Mg2+, Na+, and K+ data are retained for soil solutions. For this report, analyses focus on
surface water ANC, because this parameter provides the most integrated indicator of system response
to continued exposure to acidic deposition.
236
-------
9.3.3.6.1 Prediction of current conditions -
The distribution of current surface water ANC values projected for the M-APP Region using the
Reuss model is illustrated in Figure 9-21. The ANC values for M-APP streams, from National Stream
Survey data (Kaufmann et al., 1988) and from initial model calibrations, are listed in Table 9-20. An
obvious feature of the model projections is an extremely tight clustering of simulated ANC, with a
mean value of-21 fieq/L (standard deviation = 8/*eq/L). Clustered values of predicted ANC values
have been obseived in all regions for which Reuss model runs have been completed. The primary
difference between projections for the M-APP and those for the NE and SBRP regions is the mean
ANC values observed in each region. For the M-APP, the soils appear to be more highly acidic than
those in either of the other two regions, resulting in significantly lower predicted surface water ANC
and pH values.
As concluded for the other regions, model results are consistent with the hypothesis that soil
exchange reactions can buffer soil water and surface water ANG values, but that in acid soils such as
those that occur in the DDRP watersheds, ANC values are buffered within a low range of values. The
low, tightly clustered ANC values simulated by the Reuss model suggest that exchange is not the
primary mechanism controlling the compositions of surface waters in DDRP study regions. The
buffering provided by the soils is effective in an ANC range that is lower than the mean values
observed. Figure 9-22 illustrates the relationship between observed and projected ANC values.
Given the difference between measured and simulated values of ANC, it seems obvious that
other mechanisms in addition to exchange play a significant role in the control of surface water base
cations and ANC. Uptake of cations by aggrading vegetation represents a sink for cations, and would
be expected to deplete soil base cations and to depress surface water ANC values below those
computed by the model. The presumed base cation/ANC source, and the presumed major process
altering cation balance and ANC in the M-APP watersheds, is primary mineral weathering. Release of
base cations and ANC through weathering reactions, as discussed in Section 3.4 of Church et al.
(1989), can increase surface water ANC to values well above the limit apparently imposed by soil
exchange processes.
According to the knowledge gained from modeling activities in the three DDRP regions,
mineral weathering is apparently the major ANC generating process in those systems with ANC values
in excess of about 50 /*eq/L For systems with lower ANC values, both mineral weathering and soil
exchange processes could be influencing observed surface water chemistry. Available data, however,
preclude any reliable apportionment of the contribution from each of these two processes to the total
base cation and ANC levels in surface waters considered by DDRP.
The implications of these findings are significant in terms of projected future changes in
surface water chemistry. If mineral weathering is, in fact, regulating ANC levels in those systems with
ANC greater than 50 to 100 ^eo/L, then weathering will continue to provide base cations to neutralize
at least a significant fraction of deposition inputs. Because exchange represents a modest or even
negligible net source of base cations in these systems, depletion of the soil reservoir of exchangeable
237
-------
Reuss Mode
Year 0
Current Deposition
O
0.0
-50
-40 -30 -20 -10
SIMULATED ANC faeq/L)
o
10
Figure 9-21. Cumulative distribution of year 0 calibrated values of ANC, for the target
population of DDRP watersheds in the M-APP Region, based on simulations using
the Reuss model with Typical Year deposition estimates.
238
-------
Table 9-20. Summary Statistics for Current ANC in DDRP Target Population Stream Reaches
in the M-APP Region8
Meanb
Standard Deviation
Min
25th percentile
Median
75th percentile
Max
0 National Stream Survey Data from
DDRP Population
Estimates
87.8
58.7
-9.6
48.5
86.8
48.5
199.6
Kaufmann et al., 1988. Units are^eq/L.
Reuss Model
Estimates
-21.0
8.4
-33.7
-28.6
-20.9
-15.7
5.1
ANC for surface water.
239
-------
cr
0)
=1
200
150-
100-
Q
£ 50 H
ID
c/5
o
-50
1:1
• I
«
-50
0 50 100 150
OBSERVED ANC Qieq/L)
200
Figure 9-22. Comparison of measured values of ANC for DDRP watersheds in the M-APP
Region and values simulated using the modified Reuss model with Typical Year
deposition estimates. Measured values are from Kaufmann et al. (1988).
240
-------
base cations will occur more slowly than in otherwise comparable systems in which exchange is a
major or dominant net source of bases. The potential rate of decline in ANC attributable to depletion
of soil base cations in these systems will tend to vary in direct proportion to the relative importance of
exchange as a net cation source to runoff, and as such will be inversely related to initial ANC. In the
long term, as long as rates of mineral weathering exceed rates of deposition, ANC will remain above
zero. In contrast, in systems with ANC values less than about 50 /ieq/L, rates of sulfate deposition
probably exceed rates of primary mineral weathering, such that soil exchange processes are contrib-
uting to the maintenance of ANC. Some of the low ANC systems are probably experiencing acceler-
ated base cation leaching in. response to ongoing increases in sulfate leaching (Section 9.2). These
systems could experience significant ANC decline in the future, as sulfate mediated leaching continues
to increase and as the exchange pool is gradually depleted. Results from the Reuss model are not
sufficient, in and of themselves, to effectively identify those systems at highest risk for such changes.
9.3.3.6.2 Projected future conditions --
In order to assess the po'tential magnitude of changes in ANC that might occur in the M-APP
watersheds, and to characterize the time frame over which such changes might occur, simulations
were run for a 100-year period, using the Reuss model with a variety of deposition scenarios. The
mass balance component of the model tracks the changes in the base cation pool on soil exchange
sites through time. For these simulations, precipitation quantity (cm/yr) and the deposition fluxes were
used to specify the total loadings of ions to the soil. Annual time steps were used for these simula-
tions. For the M-APP, model simulations were run for 100 years, with results reported for years 20, 50,
and 100.
Projected time-dependent changes in ANC values for the population-weighted results are illus-
trated in Figure 9-23 and summary statistics are listed in Table 9-21. The Reuss model considers only
the effects of the soil cation exchange process in making these projections. Mineral weathering reac-
tions would supply base cations to the soil system and would, in general, diminish and delay the
response of these systems to the effects of acidic deposition. At 20 and 50 years, most systems in
the M-APP are projected to experience minimal change in ANC. The soil buffering capacity in these
systems is sufficient to moderate the effects of acidic deposition over these time scales. At current
deposition, none of the systems modeled in this region are projected to experience losses of ANC of
more than 20 /^eq/L within the 50-year time frame.
For longer time frames, however, the changes in base cations and in ANC become much more
pronounced, as soil exchange pools become depleted of base cations (Figure 9-23). About a third of
the watersheds in the region are projected to experience minimal changes (< -25 jueq/L) over
the 100-year time frame, but others are projected to experience^ much larger change in ANC (median
change of-73/^eq/L) with a maximum change of -237 ^eq/L Examination of the results (Table 9-21)
shows that projected changes in the ANC values through time are not linear, but accelerate dramatic-
ally between years 50 and 100, as the buffer capacity of soils is depleted. Soil response to acidic
deposition is analogous to a buffer being titrated by acidic deposition. When the system is above the
inflection point of the titration curve, the response of a soil to a unit loading of acidic deposition will be
a gradual, almost linear, decline in projected ANC. Once the system reaches the inflection point,
241
-------
o
DC
O
0.
o
DC
Q_
LLI
P
_J
ID
^
ID
O
1.0
0.8 -
YEAR 20
YEAR 50
YEAR 100
-250 -200 -150 -100 -50 0 - 50
PROJECTED CHANGE FROM CURRENT ANC (|ieq/L)
Figure 9-23. Cumulative distributions of projected changes in ANC at 20, 50, and 100
years, for the target population of DDRP M-APP watersheds, based on Reuss
model simulations with Typical Year deposition estimates.
242
-------
Table 9-21. Summary Statistics for Projected Changes in Surface Water ANC Values In
Streams in the DDRP Target Population in the M-APP Region at 20, 50, and 100
Years for Three Deposition Scenarios3
Scenario A
Current
Scenario B
50% Decrease
Scenario C
50%/20% Decrease
Current Mean
-21.0
-21.0
0 Units are/ieq/L
All values, except for current year estimates, are presented as changes from current status.
-21.0
Year 20b
Mean
Std. Dev.
Median
Max (negative)
Year 50b
Mean
Std. Dev.
Median
Max
Year 100b
Mean
Std. Dev.
Median
Max
-0.4
1.0
-0.3
-2.8
-3.4
3.8
-3.3
-19.1
-73.0
61.3
-72.5
-236.9
+9.8
2.6
9.6
5.6,
9.2
3.1
9.3
3.5
4.2
7.5
6.2
-17.3
8.9
2.4
8.7
5.0
2.5
2.5
3.2
-2.7
-14.8
20.2
-7.8
-93.7
243
-------
however, the rate of decline in ANC dramatically accelerates because the buffer capacity of the system
has been depleted.
Results of these simulations have major implications for soils and surface waters in the M-APP
Region. Small, gradual changes in surface water ANC in the short term do not preclude the possibility
that more rapid dramatic changes can occur in the future. This is especially true for systems with low
current ANC. Rates of change in system response can increase with time, although changes will be
moderated to some extent by mineral weathering.
9.3.3.6.3 Ramped deposition scenarios ~
The last group of simulations addresses the effects that changes in the level of sulfur deposition
have on projected future changes in ANC. As described in Section 5.6, three deposition scenarios
were considered for M-APP simulations. The first scenario is continued deposition at current levels.
The second scenario (Scenario B) ramps deposition linearly from current levels, starting at year 5, to
50 percent of current deposition by year 15. Deposition is then maintained at this lower level for the
remainder of the simulation. The third scenario (C) entails a similar ramping of deposition to 50
percent of current levels between years 5 and 15, but then gradually increases the fluxes of acidic
sulfur deposition to 80 percent of current values between years 15 and 50. After year 50, deposition is
maintained at 80 percent of current levels.
Differences between the projections made using ramped and constant deposition are summa-
rized in Table 9-21. Differences among the scenarios are substantial at the 20-year point. Since the
simulations do not include any source of base cations for the soil exchange complex, except to the
extent that soils might retain cations from deposition, the projected increases in ANC for the waters
evolving from these soils is almost entirely a mobile anion effect. That is to say, the simple act of
reducing by half the levels of acidic sulfate deposition and flux through the soil results in a decrease in
acid cation transport and an increase in projected surface water ANC values.
The median changes projected for ANC at year 50 using deposition scenario B suggest that the
ANC of surface waters in the M-APP would remain comparable to, or slightly higher than present
conditions. For the 50 percent reduction in deposition, projected surface water chemistry remains
almost unchanged during the 30-year interval. For deposition scenario C, which embodies both the
initial decline and a subsequent increase in deposition, the model projections suggest that the ANC of
runoff will increase, then gradually decline. In the absence of a mineral weathering component,
projected ANC will ultimately decline if sulfur deposition exceeds base cation inputs, until the
exchange complex is depleted. At that time, the soil system will attain a new dynamic equilibrium, in
which the soil solution and runoff will have low pH and ANC because the exchange complex can no
longer neutralize incoming acidity.
After 100 years, the differences in the medians among the three deposition scenarios are quite
marked. Simulations using the constant deposition scenario project major reductions in projected
surface water ANC, with a median decrease of more than 70 ,weq/L Scenario B projects surface water
244
-------
ANC that is little changed from present conditions; scenario C projects small declines in median ANC
(-8 fteq/L compared to -73 /*eq/L at constant deposition).
9.3.3.6.4 Comparison of results from the M-APP, NE, and SBRP -
Comparison of the results obtained using the Reuss model in the three DDRP regions indicates
both similarities and differences among the regions. Table 9-22 summarizes current surface water
chemistries projected by the model for each of the three regions. Soils in each region behave as
strong buffers, regulating the surface water ANC values to within narrow ranges; in all cases, median
ANC values are substantially lower than measured ANCs. These observations again suggest the key
role mineral weathering has in producing ANC for surface waters, even in low ANC systems that are
susceptible to the effects of acidic deposition. The predicted median ANC in the M-APP Region is
considerably lower than that predicted for systems in the other two regions. This suggests that the
soils in this region either are naturally more acidic than those in the other regions or that a substantial
fraction of their buffering capacity has been depleted as a result of long-term exposure to acidic
deposition. Other factors might also contribute to the lower predicted equilibrium ANC values for soil
leachate in soils of the M-APP Region, but reasons are unclear.
The soils In the three regions respond quite differently to continued exposure to acidic depo-
sition (Table 9-23). Soils in the NE Region are projected to undergo the largest changes during the
first 20- and 50-year simulations. At present levels of deposition, soils in the NE appear to be more
susceptible to short-term changes in leachate ANC than are the soils in the other two regions. This
conclusion is, in a sense, counter-intuitive because soils in the NE tend to have higher levels of base
saturation than the soils in the M-APP or SBRP (see Table 9-9). The soils in the NE, however, are also
younger than those in the M-APP and SBRP, and as a result, tend to have less clay-size material.
Mean changes in the surface water ANC values in the NE decline almost linearly through time. This is
in marked contrast to the behavior of soils in the other two areas, in which the near-term changes in
projected surface water quality are minimal, but rates of decline then substantially accelerate as the
buffering capacity is depleted. This effect is especially pronounced in the M-APP, and thus is con-
sistent with the hypothesis that long-term exposure to acidic deposition in this region is significantly
depleting the buffering potential of the soils.
9.3.3.7 Conclusion
Several conclusions can be drawn from the observations made using the Reuss model and the
projected changes to surface water ANC values along the eastern United States.
• For watersheds with surface waters currently exhibiting ANC values in excess of 50 to 100
/teq/L, ANC values projected by an exchange model appear unrelated to observed ANC,
suggesting that mineral weathering is the dominant watershed process controlling ANC.
• For watersheds that are in steady state with regard to sulfate adsorption and exhibit surface
waters with ANC values of less than 50 ^eq/L, soil exchange processes may be contributing
to the observed ANC. In most systems, the observed levels are probably controlled by a
245
-------
Table 9-22. Summary Statistics for Projections of Current ANC, Predicted Using the Reuss
Model with Typical Year Deposition Estimates for the NE, M-APP, and SBRPa
NE
M-APP
SBRP
Mean
Std. Dev.
Min
25th Percentile
Median
75th Percentile
Max
8.3
19.0
-46.8
-1.8
7.4
20.6
67.1
-21.0
8.4
-33.7
-28.6
-20.9
-15.7
5.1
2.2
6.1
-14.1
-1.5
2.3
4.7
2.3
Units
246
-------
Table 9-23. Projected Changes in ANC at 20, 50, and 100 Years, Based on Reuss Model
Analyses Using Typical Year Deposition Estimates for the NE, M-APP, and SBRPa
NE
M-APP
SBRP
Current
Mean 8.3
Changes at Year 20
Mean -6.4
Std. Dev. 18.0
Median -2.0
Max -118.3
Changes at Year 50
Mean -16.1
Std. Dev. 26.4
Median -6.0
Max -138.8
Changes at Year 100
Mean -43.1
Std. Dev. 51.5
Median -22.0
Max -231.7
-21.0
-0.4
1.0
-0.3
-2.8
-3.4
7.8
-3.3
-19.1
-73.0
61.3
-72.5
-236.9
2.2
-1.2
0.7
-2.2
-2.2
-3.7
2.0
-4.0
-7.4
-23.0
16.2
-18.5
-58.8
1 Units are /ieq/L
247
-------
combination of cation exchange and mineral weathering, with a relatively larger role for
exchange in systems with the lowest ANC.
• Modelled reductions in acidic sulfate deposition in the M-APP result in significant projected
improvements in surface water ANC. The improvement is primarily derived from a mobile
anion effect, and is not attributable to any increase in soil base cation content.
9.4 SUMMARY
The Level II analyses indicate that both sulfate adsorption and cation exchange by soils have
been and will continue to be important processes delaying surface water acidification resulting from
acidic deposition in the M-APP Region. Sulfate adsorption has maintained stream water sulfate
concentrations in the region at levels well below steady state, and model projections indicate that
adsorption will continue, albeit with steadily decreasing efficiency, for a few more decades. The
sulfate status of streams in the region is in transition; model projections indicate, and monitoring data
confirm, that sulfate concentrations of streams in the region are increasing by approximately 1-2
,aeq/L/yr (Table 9-2). Modelled retention in the base year for simulations (1986) was 26 percent,
somewhat lower than the median computed retention of 44 percent, but is projected to decrease to a
median of 11 percent within 20 years, and to only 3 percent in the next 50 years. At the present time,
14 percent of M-APP watersheds are at sulfur steady state; that percentage is projected to increase to
36, 58, and 93 percent in 20, 50, and 100 years, respectively (Tables 9-4, 9-5). Median time to sulfur
steady state for the M-APP watersheds is 35 years, compared to 0 years in the NE (i.e., at least 50
percent of NE systems are presently at or above steady state), but is considerably shorter than the 61
year median for streams in the SBRP.
Because sulfur deposition to M-APP watersheds is much higher than in the NE or SBRP, com-
puted steady-state sulfate concentrations in the M-APP are, on average, much higher than in the other
regions. Sulfate concentrations in DDRP target population watersheds in the M-APP are already
higher than in NE or SBRP systems (median = 153^eq/L in the M-APP, 105,aeq/L in the NE, and 25
fieq/L in the SBRP), but at steady state, the median sulfate concentration in the M-APP (215y«eq/L) will
be almost twice that of lakes in the NE (111 ^eq/L) or SBRP streams (120/leq/L) (Table 7-2).
Although, on average, sulfur retention in M-APP streams is considerably lower than in the SBRP, the
increase in sulfate concentration as systems come to steady state is larger (median and maximum
change of 107 and 248 ^eq/L, compared to 83 and 167 ^eq/L in the SBRP).
If sulfur deposition is reduced in the M-APP, there will, of course, be proportional decreases in
steady-state sulfate concentration; the rate of response of sulfate concentration depends on the extent
and timing of reductions in deposition and on the reversibility of adsorption. In simulations assuming
complete reversibility, for most M-APP watersheds (all but those systems presently at steady state),
there will be an initial lag period if deposition is reduced, during which time the stream water sulfate
concentration is projected to continue to increase. This lag results from desorption of sulfate presently
retained on upper soil horizons. Such lags are projected to last as long as 50 years, with projected
sulfate concentrations increasing as much as 60 fieq/L (Figure 9-11). In all cases, and at any point in
time during such a lag phase, however, the projected sulfate concentrations are lower than they would
be if deposition were maintained at current levels.
248
-------
The projections for sulfur relate only to the status of sulfur in soils and runoff; time to steady
state for sulfur is not directly linked to time for systems to reach zero ANC or any other threshold
value. Although changes in sulfate mobility are a principal control on changes in both base cation
mobility and ANC, other processes also influence the rate and extent of change in cation composition
and ANC of surface waters. Surface waters can reach an ANC of zero (or any other threshold value)
before, concurrently with, or after sulfur comes to steady state.
Because of the high sulfur loadings and high sulfate concentrations in M-APP watersheds, the
potential for base cation leaching from the exchange complex, and for depletion of the labile cation
pool, is high. The Level II base cation exchange analyses represent in most respects a worst-case
analysis (i.e., weathering rate assumed to be zero, sulfate assumed to be completely mobile), intended
to place an upper bound on the potential rate of cation depletion from the soil and of related changes
in the ANC of runoff.
Most M-APP soils have formed from sedimentary bedrock, and represent the poorly weatherable
remnants of one or more previous weathering episodes. Soils in DDRP watersheds in the region,
because of the parent material mineralogy and extensive historic weathering, are naturally base cation
deficient and usually strongly acidic. Soil analyses confirm the low pH and base status of most soils
on watersheds in the target population. Model simulations for the M-APP Region, as in previous DDRP
analyses for the NE and SBRP, suggest that exchange processes can be a significant source of base
cations to buffer surface water chemistry, but that the role of exchange is probably most important in
those systems with low surface water ANC (< 100 ^eq/L). Weathering probably dominates cation
supply to systems with higher ANC.
Although the two base cation models focus on different aspects of the soil system (soil pH and
base saturation in the case of the Bloom-Grigal model; runoff chemistry for the Reuss model) results
from the two models are consistent, and reflect the importance of cation exchange as a buffer of soil
and runoff pH and ANC. Under conditions of continued deposition at current levels, the ANC of runoff
(soil leachate) is projected to decrease by only a few^eq/L during the first 50 years of simulations, but
then to decrease rapidly, with a median decrease of 70/^eq/L between years 50 and 100 of simula-
tions. Simulations using the Bloom-Grigal model project a steady decline in base saturation of soils,
as the exchange pool is depleted during the 100-year period (from 13.8 to 9.4 percent). Soil pH
changes relatively little during the first 50 years, but the decrease accelerates between years 50 and
100.
Under conditions of reduced deposition, projected changes in runoff chemistry reflect both the
depletion of the exchange buffer capacity and the salt effect. If deposition is reduced by 50 percent,
Bloom-Grigal projections indicate only small, slow future decreases in soil pH (< 0.1 units) and base
saturation (from 13.8 to 13.2 percent) during the 100-year period. For the same deposition scenario,
Reuss model projections indicate an increase in the ANC of runoff of slightly less than 10/^eq/L,
attributable to a salt effect as sulfate concentrations decline. For the intermediate deposition scenario,
soil base cations and pH are projected to initially be almost unchanged as sulfur deposition decreases,
then to deteriorate as sulfate concentrations increase in the latter part of the simulations. Reuss model
projections of ANC show an initial increase in ANC of almost 10 ^eq/L (salt effect), but ANC then
249
-------
decreases below initial values at years 50 and 100, due to a combination of increased sulfate-
mediated leaching and depletion of the pool of exchangeable base cations. Overall, results of Level II
analyses can be summarized as follows:
• Sulfate adsorption and cation exchange are both important processes that act to delay the
effects of acidic deposition in soils and surface waters of the M-APP Region.
• Both adsorption and exchange are capacity-limited processes that are being depleted.
Sulfate concentrations are currently increasing as the adsorption capacity of soils becomes
saturated, and base cations are projected to be gradually depleted from the soil under
conditions of continued deposition at current levels.
• Under a scenario of continued deposition at current leveis, a median increase in sulfate
concentrations of more than 100 ^eq/L is estimated as systems come to steady state. As a
result of the high rates of sulfate leaching in the M-APP, runoff ANC is projected to have a
small median decrease (3 ^eq/L) in the first 50 years of simulations, but an additional 70
fieq/L decline between years 50 and 100. During the same 100-year period, the efficiency
of the soil base cation buffer systems will be seriously reduced, and final base saturation is
projected to be less than 10 percent.
• Sulfate concentrations are projected to decrease, and runoff ANC to increase if deposition is
reduced by 50 percent. The net depletion of base cations from the soil exchange pool
would be virtually stopped by such a reduction in deposition.
250
-------
-------
SECTION 10
LEVEL III ANALYSES - DYNAMIC WATERSHED MODELLING
10.1 INTRODUCTION
Previous sections discuss (1) the genera! DDRP approach (Section 4), (2) sulfur retention attri-
butes of Mid-Appalachian (M-APP) watersheds (Section 7), and (3) future changes in sulfate adsorp-
tion and base cation exchange (Section 9) that might occur in watersheds over the next 100 years.
This section discusses the application of a dynamic watershed model to project future changes in
surface water chemistry, referred to as Level III analyses.
Three terms are used to describe simulations of future change:
• Predict - to estimate some current or future condition within specified confidence limits on
the basis of analytical procedures and historical or current observations.
• Forecast - to estimate the probability of some future event or condition as a result of rational
study and analysis of available data.
• Project - to estimate future possibilities based on rational study and current conditions or
trends.
The model output is intended to distinguish among these three terms and definitions. Level III
analyses are defined and intended for use as projections.
Predictions typically are made to compare different scenarios, controls, or management options.
Predictions can be performed within specified confidence limits because of previous model evalua-
tions, testing, applications, and comparisons with measured data for a variety of system types. Model
predictions of various surface water attributes are legally required for many proposed management
strategies that range, for example, from examining potential alterations of hydrologic regimes by land-
use modifications to estimating mixing zones for effluent discharges to estimating phytoplankton
response to nutrient reduction. Predictions generally are performed for short time periods (e.g., single
events, parts of a season, or a few years) and focus on before-after comparisons such as water quality
before and after wasteload reductions or plankton biomass before and after nutrient reductions.
Forecasts convey some estimate of the likelihood or probability that various conditions or events
will occur in the future. Daily weather forecasting, with associated probabilities of showers, thunder-
storms, etc., is a classic example of forecasting. This represents a short-term forecast. Weather
forecasts also are made for annual or decadal time frames. Flood forecasts can be short term (daily
or weekly), but also are made for long-term events such as the probability of 100-, 1,000-, and
1,000,000-year events (NRC, 1988).
Projections, in contrast, are not accompanied by estimates of the probability that any of the
conditions or events might occur in the future. Projections can be used as a basis for relative
comparisons among various emission or deposition scenarios. Although the probability that a
scenario will occur cannot be estimated, projections do provide a relative basis for comparing costs
251
-------
and beneficial or deleterious effects associated with different control or management strategies. This
information generally is relevant to policymakers and decisionmakers for evaluating different control
strategies. The models in the Level III analyses are being used for projecting, not forecasting, the
effects of alternative acidic deposition scenarios on future changes in surface water acid-base
chemistry.
In Level III analyses, integrated, process-oriented watershed models are used to project long-
term changes (E.e., up to 100 years) in surface water chemistry as a function of current and alternative
levels of sulfur deposition. The watershed models integrate our current understanding of how various
processes and mechanisms interact and respond to acidic deposition. These mechanisms include
soil-water interactions (including soil-water contact time), sulfur retention, base cation exchange, and
replacement of base cations through mineral weathering, and other watershed processes (e.g., vege-
tative uptake, in-lake processes, organic interactions). The present study, however, does not establish
the adequacy of the formulations that implement these processes, the mode of spatial aggregation of
data, or the calibration approaches used for long-term acidification projections.
The DDRP is an applied project that has used existing techniques and models for these
analyses (Church et al., 1989). The watershed model that we have applied to watersheds in the
M-APP Region is the Model of Acidification of Groundwater in Catchments (MAGIC). The application
of this model to achieve the objectives of the DDRP was approved by peer reviewers in accordance
with the Agency's standard competitive funding process and requirement for external review of
environmental data collection programs (Section 4.4.3). The MAGIC model has been thoroughly
described in the literature (Cosby et al., 1985 a,b,c) and has been reviewed by Jenne et al. (1989) and
by Thornton et al. (1990) in the NAPAP State-of-Science/Technology Report No. 14. The MAGIC appli-
cations for lakes in the NE and streams in the SBRP were extensively reviewed as part of the 1989
DDRP Report (Church et al., 1989).
This section presents:
• a brief description of MAGIC,
• MAGIC calibration and confirmation on Coweeta watersheds,
• MAGIC projection and uncertainty procedures,
• regional population estimation and uncertainty procedures,
• regional comparisons and uncertainties, and
• discussion and conclusions. •
Church et al. (1989) presented information on:
• dynamic watershed models used in the Level III analyses,
252
-------
• preparation of the modelling datasets (specifically identifying any differences required for
Level III analyses compared to Level I and II analyses),
• general modelling approach,
• model sensitivity analyses, and
• regional projection refinements.
The procedures that were used for the NE and SBRP (Church et al., 1989) were also followed in the
MAGIC application in the M-APP Region.
10.2 MODEL OF ACIDIFICATION OF GROUNDWATER IN CATCHMENTS (MAGIC)
MAGIC is a lumped-parameter model of intermediate complexity, originally developed to project
the long-term effects (i.e., decades to centuries) of acidic deposition on surface water chemistry. One
of the principal assumptions inherent in the model is that a minimum number of critical processes in a
watershed influence the long-term response to acidic deposition. The model simulates soil solution
chemistry and surface water chemistry to project the monthly or annual average concentrations of
surface water acid-base chemical constituents (Church et al., 1989; Thornton et al., 1990). Hydrologic
flow of water through soil layers to the receiving system is simulated using a separate hydrologic
model, TOPMODEL (Homberger et al., 1985). TOPMODEL is a topography-based, variable contribu-
ting area, catchment model adapted from the version of Beven and Kirkby (1979). The model con-
siders overland flow, macropore flow, drainage from the upper zone to the lower zone and to the
stream, and baseflow from the lower zone. TOPMODEL provides flow routing through the watershed
to MAGIC, where both equilibrium and rate-controlled expressions are used to represent geochemical
processes. Mass balances for the major cations and anions and the effects of aqueous aluminum and
organic acid species on ANC are incorporated into MAGIC. The components included in the ANC
calculation are:
2'
'].- [H+] - 3[AI3+]
ANC = [HC03-j + 2[C032-] + [OK] + [HR"'] + 2[R"']
- 2[AI(OH)2+] - [AI(OH)2+]
where the terms [HR"~] and [R"2"] indicate organic acid fractions.
The ANC simulated by MAGIC is similar to the modified Gran ANC, but does not consider
organic acid contributions as part of the DDRP application. MAGIC includes explicit watershed geo-
chemical processes of carbonic acid and aluminum chemistry, mineral weathering, anion retention and
cation exchange. It implicitly considers biochemical processes of soil nitrification, nutrient uptake and
soil respiration (Thornton et al., 1990).
MAGIC represents the watershed with two soil-layer compartments. These soil layers can be
arranged vertically or horizontally to represent the vertical or horizontal movement, respectively, of
water through the soil. A vertical configuration was used in the DDRP, and the soil compartments
253
-------
were assumed to be areally homogeneous. Annual output is the typical temporal resolution of the
model, but monthly output also can be obtained. The model requires standard meteorological inputs
(i.e., precipitation, air temperature) and deposition inputs for base cations, acid anions, ammonia, and
fluoride on a monthly or annual time scale.
MAGIC was originally formulated to be parsimonious in selecting processes for inclusion and
was intended to be used as a heuristic tool for understanding the influences of the selected processes
on surface water acidification. The spatialAemporal formulations in the model reflect the intended use
for assessment and multiscenario evaluations. It was originally developed and applied to a stream in
the Mid-Appalachian Region (White Oak Run, Shenandoah National Park, VA; Cosby et al., 1985a,b,c).
Subsequently, MAGIC has been applied to numerous individual watersheds (Cosby et al., 1986c;
Cosby et al., 1990; Jenkins et al., 1990a; Lepisto et al., 1988; Neal et al., 1986; Whitehead et al.,
1988a,b). It has also been applied in a regional manner to groups of watersheds in the United States
(Homberger et al., 1986), the United Kingdom (Jenkins et al., 1990c; Musgrove et al., 1990) and in
Norway (Cosby et al., 1987; Homberger et al., 1989; Wright et al., 1991). As described previously, it
was one of three models applied in the DDRP studies in the Northeast and Southern Blue Ridge
Province (Church et al., 1989).
The MAGIC model has been tested more than any other watershed acidification model. Projec-
tions using MAGIC have been compared to (1) diatom reconstructions of lake chemistry, (2) soil
column leaching experiments, and (3) whole-watershed acidification manipulations. Applications of
MAGIC to diatom reconstructions for four lakes (Big Moose Lake in the Adirondacks, Loch Grannoch
in Scotland, Lake Gardsjon in Sweden, and Lake Howatn in Norway) "indicate that the processes
linked in MAGIC can account for temporal trends in pH and alkalinity such as those obtained from
paleolimnological data" (Wright et al., 1986). Jenkins et al. (1990b) compared MAGIC hindcasts of
preindustrial (pre-1850) pH with paleolimnology reconstructions for six Scottish lochs. The two
approaches yielded comparable historical estimates. Sullivan et al. (1991) compared hindcasts by
MAGIC and paleoecological methods for reconstruction of pre-industrial (pre-1850) surface water
chemistry for 33 Adirondack lakes studied as part of both the DDRP (Church et al., 1989) and the
Paleoecological Investigation of Recent Lake Acidification (PIRLA II) (Charles and Smol, 1990). Both
methods indicated that acidification had occurred. The MAGIC hindcasts indicated that acidification
had occurred across much of the ANC range, whereas the paleolimnological procedures (called
"Diatom" by Sullivan et al., 1991) indicated acidification primarily for lakes with current ANC less than
50 |ieq/L. MAGIC hindcasts indicated higher pre-1850 ANC values than did Diatom. Agreement
between the methods for pre-1850 pH was very good above about pH 6, with MAGIC indicating higher
historical pH values than Diatom below that level. More recently, Cosby and Ryan (1991) have incor-
porated a triprotic representation of organic acids using the pK values of Driscoll et al. (1990) for
Adirondack lakes into these MAGIC hindcasts and have found that this yields MAGIC hindcasts of pH
highly similar to those of Diatom. Differences remain in the hindcast ANC values, probably because
the Diatom model has a "stronger physiological basis" in pH (Sullivan et al., 1990) than it does in ANC
(Cosby, pers. comm.). ,
With regard to comparisons to soil column leaching experiments "(a) formulation of the soil
chemical submodel of MAGIC predicted closely the results of simple laboratory column experiments
254
-------
when flushing effects associated with drying of soils prior to the experiment were eliminated."
"Despite...reservations the results largely substantiate the validity of the fundamental principles behind
the chemical submodel of MAGIC as applied to the Loch Dee catchment. It is particularly encouraging
that a lumped model such as this can be applied at a very detailed scale and remain generally valid"
(Grieve, 1989).
Norton et al. (In press) have used MAGIC to project possible effects of ammonium sulfate appli-
cations to West Bear Brook catchment (Maine, USA) as part of the Bear Brook Watershed Manipula-
tion Project. Early comparisons of predicted and observed effects indicate that "after one year of
treatment, the response of the stream chemistry and the response modelled by MAGIC are similar"
(Norton et al., In press). Wright et al. (1990) used MAGIC to predict the result of whole-catchment
manipulations in the Reversing Acidification in Norway (RAIN) project. "Application of the MAGIC
model to the whole-catchment experimental manipulations of the RAIN project shows that the
response of water and soil acidification to large and rapid changes in acid deposition can be pre-
dicted. The aggregation approach used in the MAGIC model to describe a complex system of spati-
ally varying chemical and biological reactions produces predicted trends in runoff chemistry that agree
satisfactorily with observed trends." 'These results reinforce other evaluations of the model obtained
by comparison with paleolimnological reconstructions of lake acidification and changes in regional
lake chemistry. Together these applications give us confidence in the use of the MAGIC model as a
robust tool for predicting future soil and water acidification following changes in acid deposition"
(Wright et al., 1990).
10.3 OPERATIONAL ASSUMPTIONS
Several operational assumptions are associated both with DDRP and MAGIC (Table 10-1).
These assumptions underlie the DDRP Level III analyses in toto. Additional MAGIC assumptions exist
beyond those made for the DDRP. These specific assumptions, summarized by Jenne et al. (1989),
are described in more detail by the authors and developers of the model (Cosby et al., 1985 a,b,c).
10.4 MODELLING DATASETS
The MAGIC model, and its hydrologic model, TOPMODEL, require:
• meteorological/deposition data,
• runoff estimates,
• basin morphometry,
• soils, and
• surface water chemistry.
-------
Table 10-1. Level III Operational Assumptions
1. Index sample water chemistry from the NSWS provides an index of chronically acidic systems
and systems with low ANC that are susceptible to acidic deposition.
2. index soil data from the DDRP Soil Survey adequately characterize watershed attributes
influencing surface water chemistry.
3. Projections of future acidification consider primarily chronic acidification. Episodic acidification is
considered in the EPA Episodic Response Project (Wigington et al., 1990).
4. Surface water acidification is a sulfur-driven process. Sulfur is assumed to be the primary
acidifying agent in acidic deposition. Eastern deciduous forests generally are nitrogen-limited
(Likens et al., 1977; Swank and Crossley, 1988) so there is low export of nitrate. In addition,
annual nitrate deposition exceeds annual ammonium deposition in the eastern United States
(Kulp, 1987) and nitrate has a slight alkalizing effect in the watershed (Lee and Schnoor, 1988).
5. The watershed processes controlling the effects of sulfur deposition on surface waters are
sulfate adsorption and desorption and base cation depletion and resupply through mineral
weathering and exchange.
6. The effects of organic acids on acid-base chemistry are constant through time and independent
of sulfate.
7. These major processes are known well enough to be incorporated into MAGIC.
8. Current watershed attributes and conditions (e.g., climate, land use, basin characteristics) will
remain relatively constant over the next 50 years.
9. Long-term projections using MAGIC are plausible and such projections are a feasible approach
for evaluating the regional, long-term effects of sulfur deposition scenarios on surface water
chemistry.
10. "Typical" year projections are not intended to represent future forecasts of water chemistry but
rather to provide a common basis for comparisons among deposition scenarios to assess
potential changes in surface water chemistry.
11. Acidification is reversible and the processes in MAGIC are adequate for describing both
chemical acidification and deacidification.
12. Physical and chemical processes are adequately considered in MAGIC.
13. Uncertainty calculations provide estimates of relative error for long-term comparisons among
deposition scenarios but are not absolute error estimates.
256
-------
10.4.1 Meteorological/Deposition Data
Section 5 discusses the meteorological and deposition data, with the exception of daily meteoro-
logical data, which are specific to the Level III analyses. Meteorological data for daily temperature also
are required as model simulation input for TOPMODEL. These data are measured at fewer locations
than daily precipitation. Typical meteorological year (TMY) data have been produced for 238 locations
across the United States. These locations are usually in major cities. Ten different TMY sites were
selected and matched to each DDRP site, based on geographic location and elevation. Temperatures
were adjusted to match 30-year normal temperatures for the period 1951-80. TMY temperature data
also were adjusted to closely match long-term monthly average temperatures at the TMY sites.
Hence, the monthly and daily temporal pattern for each TMY site was representative of the long-term
norm. Because temperature is elevation dependent, the TMY data were adjusted to match the annual
30-year normal temperature at a nearby site with an elevation comparable to the National Climatic
Data Center (NCDC) site assigned to a DDRP lake. The adjustment was additive based on the differ-
ence between the annual average TMY temperature and the annual 30-year normal temperature.
Watershed specific "typical year" meteorological and deposition data were used for each water-
shed in the M-APP Region. These typical year data were repeated year after year for 100 years
(Scenario A) in performing the watershed projection under current deposition levels (Figure 10-1).
Two alternative deposition scenarios also were investigated. The first alternative deposition scenario
(Scenario B) followed the temporal sequence of current deposition levels for the first 5 years of the
projection, ramped sulfur deposition down for the next 10 years to achieve a 50 percent reduction in
current deposition, and retained sulfur deposition at this reduced level for the next 85 years. For the
second alternative (Scenario C), sulfur deposition was retained at current deposition levels for the first
5 years, was reduced by 50 percent over 10 years, but then was ramped up to 20 percent below cur-
rent deposition levels over 35 years and retained at 20 percent below current deposition for the next
50 years (Figure 10-1).
10.4.2 DDRP Runoff Estimation
The DDRP study sites are not gaged, so measured estimates of runoff were unavailable. A
combination of techniques was used, therefore, to obtain estimates of annual and monthly runoff for
the M-APP watersheds.
10.4.2.1 Annual Runoff
Annual runoff was estimated for each of the 36 M-APP watersheds, as discussed in Section 5.7.
Long-term average annual runoff estimates were based on 1951-80 records. The annual runoff was
partitioned into average monthly fractions for use in calibrating the hydrologic submodel, TOPMODEL.
10.4.2.2 Monthly Runoff
Monthly runoff fractions for each M-APP DDRP site were determined using the long-term monthly
mean discharges from USGS gaging station records. Because none of the DDRP basins had gaging
257
-------
Change from 1985 (X)
.£ o
SOX decrease
20 30 40
Time, yrs.
Figure 10-1. Three deposition scenarios used in the M-APP Region. The constant (current)
deposition scenario is Scenario A. The 50 percent decrease is Scenario B
The 50/20 percent decrease is Scenario C.
258
-------
stations at the downstream nodes, nearby gaging stations in basins with similar characteristics were
assumed to have similar monthly runoff patterns.
Initially, the USGS gaging sites and DDRP sites were co-located on the M-APP runoff contour
map (Krug et al., 1990). Any USGS gaging stations outside a 50-mile radius of the cluster of all DDRP
sites were eliminated from further analysis. Monthly and annual mean discharges for the period of
record for each station were obtained from the USGS (D. Graczyk, pers. comm.) for each of the
remaining gages. The monthly flow fractions were computed as the ratio of the monthly average dis-
charges to annual discharges for the period of record. The monthly runoff fractions were plotted to
determine the magnitude of variation in monthly runoff between stations (e.g., Figure 10-2). The
difference between monthly fraction curves for the USGS stations was no greater than approximately 5
percent per month with the exception of two stations. One station was regulated upstream by mine
pumps and the other station had only nine years of record. The two stations were omitted from further
consideration. The monthly flow fractions were assigned to DDRP watersheds based on the following
criteria:
• proximity of the basin and its position within runoff contours relative to the DDRP site,
• similarity in basin topography, size and elevation to the DDRP site, and
• quality of the discharge record (reported by the USGS).
Table 10-2 shows the relationship between the DDRP sites and the USGS gaged sites, including
the calculated monthly fractions for each DDRP site and the parameters used to assign the gage to
the DDRP site.
10.4.3 Watershed and Stream Morphometrv Data
Basin, lake, and stream morphometry and characteristics are discussed in Section 5.4. These
data, obtained from the DDRP Soil Survey and the NSS for the M-APP, were used in model calibration
for each watershed in the M-APP.
10.4.4 Soil/Stream Physical and Chemical Data
The soils data, discussed in Section 5.5, were obtained from the DDRP Soil Survey, aggregated
(Section 9), and used for model calibration. Surface water chemistry data were obtained from the
NSS and have been described in detail by Kaufmann et al. (1988).
10.4.5 Other Data
Watershed data such as bedrock geology, land use, vegetative cover, estimated depth to bed-
rock, and other data also were used for calibration of the individual watersheds. These data are
discussed in Sections 5.4 and 5.5.
259
-------
Figure 10-2. Comparison for consistency among different USGS gages within the M-APP
Region for monthly flow fractions.
260
-------
Table 10-2. Monthly Runoff Fractions for Selected USGS Gaging Stations
USGS STA.
01542000 (PA)
ETSPDR"
MEAN ANNUAL
RUNOFF (cfe)
111
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
8.24 10.41 17.59 16.40 12.95 8.09 4.09 3.17 2.55 3.42 5.17 7.94
DDRP SITE
2C028069L(PA)
2C028075L(PA)
2C028070L(PA)
USGS STA.
01544500 (PA)
ETPDR"
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
19
20
19
MEAN ANNUAL
RUNOFF (cfs)
226
EQUIVALENT MONTHLY RUNOFF DEPTH pn)
JAN
1.56
1.65
1.56
FEB
1.98
2.08
1.98
MAR
3.34
3.52
3.34
APR MAY
3.12 2.46
3.28 2.59
3.12 2.46
JUN
1.54
1.62
1.54
JUL
0.78
0.82
0.78
AUG
0.60
0.63
0.60
SEP OCT
0.48 0.65
0.51 0.68
0.48 0.65
NOV
0.98
1.03
0.98
DEC
1.51
1.59
1.51
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN
7.82
FEB
9.40
MAR
18.51
APR MAY
18.70 12.35
JUN
6.20
JUL
2.98
AUG
1.73
SEP OCT
2.16 3.67
NOV
7.34
DEC
9.15
DDRP SITE
2C029002L(PA)
2C029016L(PA)
2C029020L(PA)
USGS STA.
0155000 (PA)
ETPDR8
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
21
19
22
MEAN ANNUAL
RUNOFF (cfs)
284
EQUIVALENT MONTHLY RUNOFF DEPTH On)
JAN
1.64
1.49
1.72
FEB
1.97
1.79
2.07
MAR
3.89 "
3.52
4.07
APR
3.93
3.55
4.11
MAY JUN
2.59 1 .30
2.35 1.18
2.72 1.36
JUL
0.62
0.57
0.65
AUG
0.36
0.33
0.38
SEP OCT
0.45 0.77
0.41 0.70
0.48 0.81
NOV
1.54
1.39
1.61
DEC
1.92
1.74
2.01
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN
7.43
FEB
8.52
MAR
17.62
APR
18.36
MAY JUN
11.66 6.29
JUL
3.35
AUG
2.40
SEP OCT
2.76 4.20
NOV
8.61
DEC
8.81
DDRP SITE
1D029023L(PA)
1 D029031 L(PA)
USGS STA.
01552500 (PA)
ETPDR3
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
21
21
MEAN ANNUAL
RUNOFF (cfs)
48.6
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
JAN
1.56
1.56
FEB
1.79
1.79
MAR
3.70
3.70
APR
3.85
3.85
MONTHLY
JAN
8.47
FEB
9.44
MAR
15.15
APR
15.51
MAY JUN
2.45 1.32
2.45 1.32
JUL
0.70
0.70
AUG
0.50
0.50
SEP OCT
0.58 0.88
0.58 0.88
NOV
1.81
1.81
DEC
1.85
1.85
FRACTION OF MEAN ANNUAL RUNOFF (%)
MAY JUN
10.69 5.52
JUL
3.14
AUG
2.57
SEP OCT
3.74 5.20
NOV
9.69
DEC
10.88
DDRP SITE
1D029042L(PA)
1D029043L(PA)
USGS STA.
01542000 (PA)
ETSDR8
EQUIVALENT ANNUAL
RUNOFF DEPTH (in) JAN
21
20
MEAN ANNUAL
' RUNOFF (cfs)
111
1.78
1.69
FEB
1.98
1.89
MAR
3.18
3.03
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
APR MAY JUN JUL AUG SEP OCT
3.26
3.10
2.24 1.16
2.14 1.10
0.66
0.63
0.54
0.51
0.79 1.09
0.75 1.04
NOV
2.04
1.94
DEC
2.28
2.18
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN
8.24
FEB
10.41
MAR
17.59
APR
16.40
MAY JUN
12.95 8.09
JUL
4.09
AUG
3.17
SEP OCT
2.55 3.42
NOV
5.17
DEC
7.94
DDRP SITE
2C035027L(PA)
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
25
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
2.06 2.60 4.40 4.10 3.24 2.02 1.02 0.79 0.64 0.86 1.29 1.98
(Continued)
Parameters used in selecting USGS stations:
E - Elevation
T - Topography
S - Size of watershed
P - Proximity of USGS Station to DDRP Station
D - Data good
R - Runoff contours
261
-------
Table 10-2 Continued
USGS STA.
01538000 (PA)
ETSOR*
DDRP SITE
2B036062L(PA)
MEAN ANNUAL
RUNOFF (cfe)
64.8
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
8.85 10.01 15.13 14.82 11.04 6.32 4.56 3.14 3.55 4.87 7.51 9.50
DDRP SITE
1D036011L(PA)
2B038028L(PA)
USGS STA.
01564500 (PA)
ETPDR*
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
17
16
MEAN ANNUAL
RUNOFF (cfe)
243
JAN
1.50
1.42
JAN
8.01
EQUIVALENT MONTHLY RUNOFF DEPTH On)
FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
1.85 2.57 2.48 1.88 1.07 0.78 0.53 0.60 0.83 1.28 1.62
1.75 2.42 2.34 1.77 1.01 0.73 0.50 0.57 0.78 1.20 1.52
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
13.63 18.19 16.43 10.80 6.54 2.53 2.12 1.87 3.68 6.27 8.93
EQUIVALENT ANNUAL
RUNOFF DEPTH On) JAN
17 1.36
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
FEB MAR APR MAY ° JUN JUL AUG SEP OCT NOV DEC
2.32 326 2.79 1.84 1.11 0.43 0.36 0.32 0.63 1.07 1.52
USGS STA.
01448500 (PA)
ETSOR1
DDRP SITE
2B041008L(PA)
MEAN ANNUAL
RUNOFF (cfe)
4.86
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
8.66 8.82 13.36 15.53 10.67 7.29 4.83 3.84 3.89 4.98 7.67 10.36
DDRP SITE
1D037005L(PA)
1D038017L(PA)
2B03S040L(PA)
USGS STA.
01603500 (PA)
ETPDR*
EQUIVALENT ANNUAL
RUNOFF DEPTH On)
27
27
25
MEAN ANNUAL
RUNOFF (cfe)
32.4
JAN
2.34
2.34
2.17
JAN
8.90
FEB
2.38
2.38
2.20
FEB
11.98
MAR
3.61
3.61
3.34
EQUIVALENT MONTHLY RUNOFF DEPTH
APR MAY JUN JUL AUG SEP
4.19 2.88 1.97 1.30 1.06 1.05
4.19 2.88 1.97 1.30 1.06 1.05
3.88 2.67 1.82 i.21 0.98 .097
On)
OCT
1.35
1.35
1.25
NOV
2.07
2.07
1.82
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
MAR APR MAY JUN JUL AUG SEP OCT NOV
19.71 16.50 11.70 7.06 3.46 2.57 2.55 4.13 4.11
DEC
2.80
2.80
2.59
DEC
7.34
EQUIVALENT ANNUAL
RUNOFF DEPTH On) JAN
15 1.34
EQUIVALENT MONTHLY RUNOFF DEPTH On)
FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
1.80 2.96 2.47 1.76 1.06 0.52 0.38 0.38 0.62 0.62 1.10
USGS STA.
01620500 (VA)
ETSDR1
DDRP SITE
2B047036L(VA)
2B047089L(VA)
2B047076L(VA)
MEAN ANNUAL
RUNOFF (cfe)
25.8
EQUIVALENT ANNUAL
RUNOFF DEPTH On)
11
15
15
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
9.17 11.84 17.01 14.79 10.93 7.69 2.02 2.38 2.23 4.63 8.43 8.88
EQUIVALENT MONTHLY RUNOFF DEPTH On)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
1.01 1.30 1.87 1.63 1.20 0.85 0.22 0.26 0.25 0.51 0.93 0.98
1.37 1.78 2.55 2.22 1.64 1.15 0.30 0.36 0.33 0.69 1.26 1.33
1.37 1.78 2.55 2.22 1.64 1.15 0.30 0.36 0.33 0.69 1.26 1.33
USGS STA.
01666500 (VA)
ETSPDR*
DDRP SITE
2B047066L(VA)
MEAN ANNUAL
RUNOFF (cfe)
220
EQUIVALENT ANNUAL
RUNOFF DEPTH On)
17
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
9.27 11.20 12.41 11.62 9.58 7.46 4.73 5.49 5.64 6.13 7.91 8.55
EQUIVALENT MONTHLY RUNOFF DEPTH On)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
1.58 1.90 2.11 1.98 1.63 1.27 0.80 0.93 0.96 1.04 1.34 1.45
Poramatets used in selecting USGS stations:
E - Etavation
T - Topography
S - Size of watershed
P - Proximity of USGS Station to DDRP Station
D - Daia good
R - Runoff contours
(Continued)
262
-------
Table 10-2 Continued
USGS STA.
03070500 (PA)
ETPDR8
DDRP SITE
2C041002L(PA)
MEAN ANNUAL
RUNOFF (cfe)
418
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
29
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEE MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
12.59 13.74 15.88 12.77 9.53 5.86 3.43 2.76 2.42 3.41 6.59 11.02
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
JAN FEE MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
3.65 3.98 4.61 3.70 2.76 1.70 1.00 0.80 0.70 0.99 1.91 3.20
USGS STA.
0306900 (WV)
ETPDR8
DDRP SITE
2C047007L(WV)
MEAN ANNUAL
RUNOFF (cfe)
771
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEE MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
10.91 11.86 15.58 12.71 9.04 6.12 4.82 4.16 3.21 4.59 7.22 9.79
DDRP SITE
2C041039L(WV)
2C041 040L(WV)
2C041045L(WV)
2C041051L(VW)
USGS STA.
03180500 (WV)
ETPDR8
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
31
33
33
35
MEAN ANNUAL
RUNOFF (cfe)
262
EQUIVALENT MONTHLY RUNOFF DEPTH
JAN
3.38
3.60
3.60
3.82
FEE MAR
3.68 4.83
3.91 5.14
3.91 5.14
4.15 5.45
MONTHLY FRACTION
JAN
11.03
FEE MAR
APR
3.94
4.19
4.19
4.45
MAY
2.80
2.98
2.98
3.16
JUN
1.90
2.02
2.02
2.14
OF MEAN ANNUAL
APR
13.82 18.08 13.89
MAY
9.31
JUN
5.24
JUL
1.49
1.59
1.59
1.69
AUG
1.29
1.37
1.37
1.46
SEP
1.00
1.06
1.06
1.12
(in)
OCT
1.42
1.51
1.51
1.61
NOV
2.24
2.38
2.38
2.53
DEC
3.03
3.23
3.23
3.43
RUNOFF (%)
JUL'
3.08
AUG
2.63
SEP
2.01
OCT
3.50
NOV
7.09
DEC
10.33
DDRP SITE
2C047010L(WV)
2B047032L(WV)
USGS STA.
03050500 (WV)
ETPDR8
EQUIVALENT ANNUAL
RUNOFF DEPTH (in)
27
19
MEAN ANNUAL
RUNOFF (cfe)
516
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
JAN FEE MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
2.98 3.73 4.88 3.75 2.51 1.42 0.83 0.71 0.54 0.94 1.91 2.79
2.09 2.63 3.44 2.64 1.77 1.00 0.59 0.50 0.38 0.66 1.35 1.96
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEE MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
12.02 14.43 16.56 12.64 8.92 5.62 3.61 3.11 2.08 3.17 6.72 11.11
EQUIVALENT ANNUAL
RUNOFF DEPTH (in) JAN
30 3.61
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
FEE MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
4.33 4.97 3.79 2.68 1.69 1.08 0.93 0.62 0.95 2.02 3.33
USGS STA.
03186500 (WV)
ETPDR8
MEAN ANNUAL
RUNOFF (cfe)
262
EQUIVALENT ANNUAL
MONTHLY FRACTION OF MEAN ANNUAL RUNOFF (%)
JAN FEB MAR APR MAY JUN JUL AUG SEP OCT NOV DEC
11.06 12.98 16.57 12.46 8.57 5.13 4.81 4.19 2.26 4.24 7.47 10.26
EQUIVALENT MONTHLY RUNOFF DEPTH (in)
DDRP SITE
2C046033L(WV)
2C046034L(WV)
2C046005L(WV)
2C046041L(WV)
2C046050L(WV)
2C057004L(WV)
RUNOFF DEPTH (in) JAN
34
37
21
30
24
22
3.76
4.09
2.32
3.32
2.65
2.43
FEB
4.41
4.80
2.73
3.89
3.12
2.86
MAR
5.63
6.13
3.48
4.97
3.98
3.65
APR
4.24
4.61
2.62
3.74
2.99
2.74
MAY
2.91
3.17
1.80
2.57
2.06
1.89
JUN JUL AUG
1.74
1.90
1.08
1.54
1.23
1.13
.64 1.42
.78 1.55
.01 0.88
.44 1.26
.15 1.01
.06 0.92
SEP
0.77
0.84
0.47
0.68
0.54
0.50
OCT
1.44
1.57
0.89
1.27
1.02
0.93
NOV DEC
2.54 3.49
2.76 3.80
1.57 2.15
2.24 3.08
1.79 2.46
1.64 2.26
Parameters used in selecting USGS stations:
E - Elevation
T - Topography
S - Size of watershed
P - Proximity of USGS Station to DDRP Station
D - Data good
R - Runoff contours
263
-------
10.5 CALIBRATION OF MAGIC
10.5.1 Calibration Procedure
MAGIC represents the horizontal dimension of the watershed as a homogeneous unit with no
subcatchments and the vertical dimension as two soil compartments. Watershed data for MAGIC were
lumped or aggregated to provide average or weighted average values for each of the soil layers. The
top soil compartment represented the mass-weighted average conditions of the A and B master hori-
zons. The lower soil compartment represented the mass-weighted average conditions in the C master
horizon.
Projecting long-term effects of acidic deposition on surface water chemistry using MAGIC
involves coupling MAGIC with TOPMODEL (Cosby et al., 1985a,b,c). Both models were calibrated
using an optimization procedure that selected parameters so that the difference between the observed
and predicted measurements was minimized. The calibration exercise was a three-step process. The
first step was to specify the model inputs such as precipitation, deposition (both wet and dry), an
estimate of historical inputs for the long-term chemical model, and fixed parameters or parameters
whose values correspond directly to (or can be computed directly from) field measurements, e.g.,
topographic variables such as slope, aspect, area. This approach, in effect, assigns all the uncertainty
associated with sampling, aggregation, and intrinsic variability to the "adjustable" parameters. The
adjustable parameters are those that are calibrated or scaled to match observed field measurements.
The second step was to select optimal values for the adjustable parameters. The optimization
method of Rosenbrock (1960) was employed. Optimal values were determined by minimizing a loss
function defined by the sum of squared errors between simulated and observed values of system state
variables. Different loss functions were used for the hydrologic and chemical models. The hydrologic
model used monthly flow estimates. The chemical model used available stream chemistry records
(i.e., weekly, monthly observations) and observed soil chemistry.
The final step was to assess the structural adequacy of the model in reproducing the observed
behavior of the criterion variables and parameter identifiability or the uniqueness of the set of optim-
ized parameters. Structural adequacy was assessed by examining the mean error in simulated values
of observed state variables for those variables used in the calibration procedure, as well as for an
additional state variable that was not used during calibration. Parameter identifiability was assessed
using approximate estimation error variances for the optimized parameters (Bard, 1974). Cosby et al.
(1989) have presented additional information on the MAGIC calibration process.
10.5.2 Calibration/Confirmation Studies
10.5.2.1 Study Watersheds
We used data from Watersheds #34 and #36 of the Coweeta Hydrologic Laboratory for calibra-
tion and confirmation studies. The USDA Forest Service Coweeta Hydrologic Laboratory, Otto, North
Carolina, encompasses about 2,184 ha (5,400 ac) and contains over 50 experimental watersheds.
264
-------
Topography at Coweeta is steep, with elevations ranging from 686 to 1,600 m and average side
slopes of 50 percent (Swank and Crossley, 1988). The Coweeta regolith is deeply weathered and
averages about 7 m in depth, which varies dramatically throughout the basin. Soils generally occur
within two orders - fully developed Ultisols and immature Inceptisols. Meteorological, deposition,
hydrologic, and stream water chemistry data are monitored at several locations throughout the basin.
Coweeta Watershed 36 is located at a higher elevation in the Coweeta Experimental Forest than
Coweeta Watershed 34 and has different hydrologic and chemical characteristics (Table 10-3). The
Forest Service has compiled a continuous 12-year hydrologic and stream chemistry record on
Coweeta Watershed 36, from June 1973 through May 1985. Data are generally available for the same
period for Coweeta Watershed 34, with the exception of stream chemistry from June 1975 through
May 1980, for which period data are unavailable. These records for the two watersheds were
partitioned into calibration and confirmation data sets to evaluate longer simulation periods, using the
MAGIC model, than were possible with the data sets available for northeastern lakes (see Church et
al., 1989). The periods were:
Watershed
Calibration
Confirmation
Watershed 36
Watershed 34
June 73 - May 79
June 82 - May 85
June 79 - May 85
June 73 - May 75
Nov. 80 - May 82
The confirmation period for Watershed 36 represented a forecast, whereas the confirmation period for
Watershed 34 represented a hindcast.
In the calibration of Coweeta watersheds 34 and 36, TOPMODEL used daily stream flow
volumes. This was a departure from the standard DDRP modelling protocol (Church et al., 1989) in
which monthly flow fractions were used in the calibration process. This should not introduce any
significant inconsistency, however, inasmuch as Coweeta has no snowpack or overland flow, and thus
we can reliably estimate (on either a daily or monthly basis) that all atmospheric deposition interacts
with one or both soil horizons (J. Cosby, pers. comm.)
10.5.2.2 Results
Calibration/confirmation results were presented previously for the intensively studied watersheds
in the NE—Woods Lake, Panther Lake, and Clear Pond (Church et al., 1989). MAGIC was developed
for southeastern streams and calibration exercises on these streams have been described elsewhere
(Cosby et al., 1985a,b,c).
The approach used to calibrate MAGIC on the two Coweeta watersheds was similar to the DDRP
calibration protocol used for the target population projections. Average annual hydrologic and
chemical conditions were computed for the multiple years of calibration data. The MAGIC model was
then calibrated to these average annual conditions, rather than the multiple years of calibration data,
to represent the DDRP approach of projecting "typical" year conditions. This approach permitted an
265
-------
Table 10-3. Inputs, Watershed Characteristics, and Surface Water Chemistries for Coweeta
Watersheds
SITE CHARACTERISTIC8
COWEETA
36
COWEETA
34
ATMOSPHERIC
Avg. Annual Precipitation (cm)
Volume Weighted pH
Sulfate Deposition (keq/ha-yr)
Nitrate Deposition1 (keq/ha-yr)
WATERSHED CHARACTERISTICS
Basin Area (km2)
Relief (m)
Depth to Bedrock (m)
Dominant Soil
Estimated SO4'2 Sorption
Vegetative Cover
SURFACE WATER CHEMISTRY
Average pH
Average ANC fueq/L)
Average SO4~Z (aeq/L)
Average NO3"4 (weq/L)
Average Annual Flow (cm)
222
4.6
0.7
0.2
0.5
520
5
Udults
high
N Hardwood-Oak-Hickory
6.5
55
19
1
168
201
4.6
0.7
0.2
0.4
330
deep
Udults
high
Oak-Hickory
6.8
85
9
0
118
0 Swank and Crossley, 1988
266
-------
evaluation of the typical year conditions used in DDRP regional calibrations and made use of the
capability of the MAGIC model for predicting longer term records of stream chemistry. Although these
projections do not confirm MAGIC'S capability for 100-year projections, they do permit a comparison
of long-term field observations with DDRP "calibrated" values and an assessment of the ability of the
MAGIC model to simulate a multi-year (< 15 yr) period of record.
The root mean squared error (RMSE) is based on differences between simulated and observed
values over the period of interest. The magnitude of the RMSE is a function of (1) the ability to
correctly simulate the mean value of the observations (i.e., lack of bias in the model), and (2) the
ability to correctly simulate the variation about the mean value (i.e., ability to reproduce seasonal
variation or long-term trends). For an unbiased model, a comparison of the RMSE to the standard
deviation of the observed data can be useful in determining the amount of seasonal or long-term
variation that is reproduced (or missed) by the simulations.
MAGIC was calibrated to the averages of the observations for the watersheds. Therefore, we
presume that the MAGIC results are unbiased with regard to the means of the observed values. We
do not expect a priori, however, to capture the variation about the mean values, because the
information on variability was not used in the calibration procedure. Therefore, a value of the RMSE
approximately equal to the standard deviation of the observations is considered acceptable in terms of
model behavior. If, in fact, the RMSE is less than the standard deviation, the model simulations are
reproducing some of the dynamic behavior of the system (i.e., the seasonal variations or long-term
trends). The calibration/confirmation results, expressed as RMSEs for Watersheds 34 and 36 are
shown in Table 10-4. For most of the variables shown in Table 10-4, the RMSEs are less than, or
approximately equal to, the standard deviations of the observations.
There was generally good agreement between the predicted versus the observed values for both
flow and calculated ANC for Watershed 36 over the 12-year period from 1973 through 1985 (Figure
10-3). The RMSE for flow was about 3 cm/mo, whereas the RMSE for ANC was about 8 ^eq/L during
the calibration period and 9 /ueq/L during the confirmation period (Table 10-4). Predicted ANC
concentrations generally were less than observed values during the calibration period and greater than
observed values during the confirmation period (Figure 10-3). In the MAGIC model, ANC is a
calculated, not a calibrated, variable. Calcium, a base cation that is included in the calibration
process, exhibited a pattern similar to that of ANC, with generally good agreement between simulated
and observed values (Figure 10-4). Simulated calcium concentrations also were slightly greater than
observed values during the confirmation period. Simulated sulfate concentrations were more variable
than calcium and were not as well correlated with observed values (Figure 10-4). The temporal
correspondence between simulated and observed values, including the slight increasing trend in
observed sulfate concentrations, was represented but there were differences in magnitude between
predicted and observed sulfate concentrations. An increasing trend in stream sulfate concentration of
about 0.7 fteq/L/yr has been observed in Coweeta Watershed 36 (Waide and Swank, 1987; J. Waide,
pers. comm.). The calibrated MAGIC trend is less than 2/u.eq/L/yr, which is expected because MAGIC
was calibrated on a single composite average annual constituent value for each calibration variable, as
per the standard DDRP protocols (Section 10.5; also, Church et al., 1989). The RMSEs for the
individual chemical species generally were less than 5 /ueq/L throughout both calibration and
267
-------
Table 10-4. RMSE Values Based on Volume-Weighted Monthly Averages of Simulated and
Observed Quantities for Coweeta Watersheds over a 12-Year Period of Record
Coweeta Watershed 36
All data
Ca
Mg
Na
K
NH4
SO4
Cl
NO3
SBCd
SAA8
Calkf
Flow8
Sim.
3.4°
2.7
3.2
1.2
1.2
4.7
4.7
1.7
8.9
8.0
8.4
3.3
Obs
4.4
2.9
3.8
1.0
.2
4.6
1.6
1.0
10.9
5.3
11.4
9.8
Calib.
Sim.
3.7
2.7
3.4
1.2
1.1
4.8
4.7
1.7
9.3
7.9
8.1
3.3
period3
Obs.
4.7
3.3
4.1
1.1
.2
3.8
1.5
1.1
11.7
4.4
10.9
10.6
Valid periodb
Sim.
3.1
2.7
2.9
1.2
1.3
4.7
4.8
1.7
8.5
8.2
8.7
3.2
Obs.
3.7
2.2
3.5
.8
.1
3.9
1.7
.8
9.3
4.8
9.2
8.9
Coweeta Watershed 34
Ca
Mg
Na
K
NH4
SO4
Cl
NO3
SBC
SAA
Calk
Flow
2
2
3
1
3
2
8
4
7
2
All
Sim
.7
.5
.3
.0
.5
.0
.0
.6
.2
.2
.7
.0
data
. Obs.
3.7
2.8
3.5
1.0
.1
2.8
1.2
.2
9.7
3.5
9.4
5.3
Calib.
Sim.
2.6
2.4
3.0
1.2
.4
2.4
1.7
.6
8.3
3.6
7.1
2.2
period*1
Obs.
3.7
2.5
3.0
1.0
.1
2.1
1.2
.2
8.8
2.7
8.4
5.3
0 Calibr. period = June 73 through May 79 (n=72 months). '
b Valid,
c Units
d Sum
0 Sum
, period = June 79 through May 85
(n=72 months)
« fteq/L (except Flow - cm/mo).
of base cations =
(Ca
of acid anions = (SO4
+ Mg + Na
+ NO3 + O
+ K).
+ F).
h
f
'
Valid
Sim.
2.8
3.1
3.8
.8
.6
2.8
2.9
.6
8.5
4.8
10.9
1.4
SBC-SAA.
period 11
Obs.
3.9
3.1
3.9
1.3
.1
3.7
1.3
.3
10.7
4.2
9.9
6.2
9 RMSE based on total
Calibr. period :
Valid, period 1
Valid, period 2
Valid period 2]
Sim.
2.7
1.9
3.3
1.0
.5
3.9
1.3
.6
7.9
4.5
4.7
1.8
Obs.
4.7
3.4
3.9
1.0
.2
2.8
1.3
.1
12.3
3.4
11.6
3.1
monthly discharge.
= June 82 through May 85 (n=36 months).
= June 73 through Apr 75 (n=
= Nov 80 through May 82 (n=
17 months).
19 months).
268
-------
120
100
COWEETA WATERSHED 36 ( simulated observed)
o
60
40
20
1974
1976
1978
1980
1982
1984
1974
1976
1978
1980
1982
1984
Year
Figure 10-3. Simulated and observed values for ANC and flow in Coweeta watershed 36
using the MAGIC model.
269
-------
50
40
30
20
10 -
0
COWEETA WATERSHED 36 ( simulated observed)
1974
1976
1978
1980
1982
1984
1974
1976
1978
1980
Year
1982
1984
Figure 10-4. Simulated and observed values for calcium and sulfate concentrations In Coweeta
Watershed 36 using the MAGIC model.
270
-------
confirmation periods. The individual base cation species typically had better precision than the
individual acid anion species, with an RMSE of about 3 fieq/L for base cations and about 5 fieq/L for
sulfate and chloride. Both potassium and nitrate, which are biologically active chemical species, had
low concentrations in the stream and low standard errors (Table 10-4). The RMSEs for sum of base
cations, however, was equivalent to that of the sum of acid anions.
Predicted versus observed flow and calculated ANC also generally agreed for Watershed 34
(Figure 10-5). The calculated ANC was somewhat lower during the 1973-75 confirmation period, but
was similar to the seasonal trends, which were consistent regardless of the calibration or confirmation
period. The RMSE for calculated ANC varied from about 5 ,«eq/L during one confirmation period to
about 10/^eq/L during the second confirmation period (Table 10-4). Simulated calcium concentra-
tions, again, exhibited a pattern similar to that of ANC (Figure 10-6).' The simulated calcium
concentrations were lower during the confirmation periods and slightly higher during the calibration
period. Simulated sulfate concentrations also were more variable than calcium and ANC concentra-
tions (Figure 10-6). Sulfate concentrations were lower than observed concentrations during the first
confirmation period, higher during the second confirmation period, and similar during the calibration
period (Figure 10-6). Simulated RMSEs for individual base cation and acid anion species in Coweeta
Watershed 34 were similar, about 3 /weq/L, for both calibration and confirmation periods, but the sum
of base cations had about twice the RMSEs of the sum of acid anions (Table 10-4).
In general, the Coweeta simulations indicate that, when calibrated according to the DDRP
protocol, MAGIC is capable of predicting seasonal trends comparable to observed data for longer term
periods of record but might not capture relatively subtle long-term trends in the record. MAGIC was
calibrated to a single composite average annual value for each calibration variable, which does not
have a trend. Although MAGIC will simulate long-term trends driven by trends in the input data, the
projected rate of change is likely to be underestimated because MAGIC did not incorporate these
trends during calibration. The DDRP protocol of calibration using typical year values captures the
general seasonal and annual pattern and should permit comparisons among deposition scenarios.
10.6 MODEL PROJECTIONS FOR MID-APPALACHIAN REGION
10.6.1 General Approach
This section discusses the general approach for performing long-term projections of the effects
of sulfur deposition on surface water chemistry over the next 100 years in the M-APP Region. The
three simulated deposition scenarios are illustrated in Figure 10-1. In the first scenario (Scenario A),
current deposition levels throughout the M-APP Region were maintained over the 100-year interval.
Changes in stream water chemistry were projected over this same 100-year period. In the second
deposition scenario (Scenario B), current deposition levels at each DDRP site were held constant for
the first 5 years, decreased by a total of 50 percent over the next 10 years, and then held constant at
this reduced deposition rate for the next 85 years. This phasing corresponded with one possible
scenario of how deposition rates might change if emissions controls were implemented (U.S. EPA,
Office of Policy, Planning and Evaluation, pers. comm.). For the third scenario (Scenario C),
271
-------
120
100
60
40
20
COWEETA WATERSHED 34 ( simulated
observed)
1974
1976
1978
1980
1982
1984
SO
40
30
E
1974
1976
1978 1980
Year
1982
1984
Figure 10-5. Simulated and observed values for ANC and flow in Coweeta watershed 34 using
the MAGIC model.
272
-------
50
40
1? 30
3
a
u 20
10
COWEETA WATERSHED 34
( _ simulated
observed)
1974
1976
1978
.1980 '
1982
1984
50
40
30
20
10
•a-
O
00
1974
1976
1978
1980
1982
1984
Year
Figure-10-6. Simulated and observed values for calcium and sulfate concentrations in
Coweeta Watershed 34 using the MAGIC model.
273
-------
deposition rates were held constant for the first 5 years, decreased by a total of 50 percent over the
next 10 years, increased to 20 percent below current deposition for the next 35 years, and then held
constant for the remaining 50 years (Figure 10-1). This scenario also corresponds with one possible
emission control strategy.
MAGIC was calibrated on individual watersheds using the data sources indicated in Section
10.5. The projected change in surface water chemistry at the NSS downstream node was simulated
for the next 100 years using the typical year meteorology and deposition data, discussed in Section
10.5. The model output represents flow-weighted annual average constituent concentrations.
Not all watersheds in the M-APP were simulated. There were some watersheds for which the
optimization and calibration criteria were not satisfied. Long-term projections for these watersheds,
therefore, were not performed (Table 10-5). In general, optimization criteria could not be achieved
either because cation inputs exceeded outputs or charge balance could not be achieved (i.e., chloride
imbalances). The results discussed in this section have been obtained by weighting the individual
watershed estimates by the appropriate inclusion probability, the weight being the inverse of the
inclusion probability (see Section 5.2, Table 5-1). Weighting by the appropriate inclusion probability is
critical for any analyses performed on these data. The emphasis in DDRP is on the relative magnitude
and direction of changes in attributes of the specified target population. The individual watershed
estimates are of interest only as they relate to the distribution of the population attributes.
A number of the stream reaches apparently were affected by inputs of roadsalts. We made this
determination the same way we did for the DDRP analyses for the NE and SBRP (see Church et al.,
1989, Section 10.5.7). To account for these inputs, we "added" salts (seasalt composition) as a point
input directly to the stream reaches during the model calibrations, exactly as we did for inland
watersheds in the NE and SBRP (Church et al., 1989).
10.6.2 Target Population for MAGIC Protections
An estimated 4,298 streams, representing 11,246 km of stream length in the target population in
the M-APP Region, were simulated using MAGIC, as compared to the DDRP total target population of
5,496 streams (15,239 stream km in the M-APP). As presented earlier (Section 5.2.7), the DDRP study
sites represent target populations that are a subset of the National Stream Survey target population
(25,715 stream reaches, 69,569 km) for NSS Subregions 1D, 2Cn, and 2Bn (Kaufmann et al., 1988).
The M-APP DDRP modelling target population reflects an exclusion of stream reaches for which
MAGIC could not satisfy the calibration criteria for the respective sample watersheds. The modelling
target population includes both disturbed and undisturbed watersheds based on chloride concentra-
tions, watersheds that had both positive and negative sulfur retention, and watersheds that had initial
NSS ANC values ranging from -14 to 198 ^eq/L. MAGIC projections for this target population are for
a 100-year period. The 100-year time frame permits an evaluation of the cumulative effects of sulfur
deposition on changes in surface water chemistry.
274
-------
Table 10-5. Watersheds and Attributes for Which Satisfactory MAGIC Calibration Was Not
Achieved
Watershed
ID
State
Attribute
2B047036
2B047066
2B047076
2C028069
2C029016
2C035027
2C047010
VA
VA
VA
PA
PA
PA
WV
Na inputs exceeded watershed outputs
S isotherms did not match observed adsorption
Na, Ca inputs exceeded watershed outputs
ANC after calibration = 103, observed ANC = -3,3 fieq/L
No successful calibrations out of 10 tries
No successful calibrations out of 10 tries
No successful calibrations °out of 10 tries
275
-------
10.6.3 Regional Calibration "
A comparison of calibrated values for ANC, pH, sulfate, and calcium plus magnesium with NSS
Phase I values is given in Figure 10-7. There is perfect correspondence between calibrated and NSS
measured values in Figure 10-7 if all the points lie on the 1:1 line. There was general agreement for
the low ANC watersheds but greater scatter between calibrated and observed ANC concentrations for
watersheds with ANC concentrations greater than 50 ^eq/L (Figure 10-7). Calibrated sulfate concen-
trations generally corresponded with NSS measured sulfate concentrations for concentrations
less than 150 ,aeq/L (Figure 10-7). There was greater scatter about the 1:1 line at the higher sulfate
concentrations. There is excellent correspondence between calibrated and observed calcium plus
magnesium concentrations over the range of DDRP watersheds (Figure 10-7). Calcium and magne-
sium are two of the variables considered during calibration optimization. There is considerable scatter
in pH over the entire range of pH values in the DDRP watersheds (Figure 10-7). Like ANC, pH is a
calculated, not a calibrated, variable.
Although there is scatter for individual watersheds, there is generally better correspondence for
the target population (Figure 10-8). The same general patterns are evident (e.g., lower calibrated than
observed ANC and pH values; greater calibrated than observed sulfate) but the effects of individual
watershed deviations are reduced at the target population level. In part, this occurs because the
population weighting factors and cumulative distribution damp some of the sample variation. The
exception is calcium plus magnesium, where almost perfect correspondence occurs on the 1:1 plot
(Figure 10-7) but larger deviations occur between calibrated and observed values for the target
population (Figure 10-8). Small errors in watersheds with larger weighting factors result in larger
deviations in the target population. In addition, once established, any deviation or offset (e.g.,
overestimate) is perpetuated through the cumulative distributions and gives an added visual
impression of difference.
There is a general bias in the calibrated DDRP M-APP target population (Table 10-6). The lower
quartile and median ANC and pH values are lower in the calibrated modelling target population than in
the comparable NSS target population. The lower quartile and median ANC concentrations for the
calibrated target population are 20 and 64 /teq/L, respectively, compared with NSS ANC concentra-
tions of 48 and 87, respectively. The lower quartile pH value for the calibrated target population is 6.2
compared with NSS values of 6.4. The median pH values are similar between the calibrated and NSS
target population. Sulfate concentrations are generally higher in the calibrated sample than in the
NSS target population.
This bias in calibrated ANC in the modelled target population results in a bias in the number and
length of acidic streams in year 0 (initial conditions) (Table 10-7). The initial calibrated percentage
(and percent length) of acidic and low ANC streams (i.e., ANC < 50 fieq/L) was 9 percent (5 percent)
and 38 percent (33 percent), respectively. For the target population modelled in the M-APP DDRP, the
percentage (and percent length) of acidic and low ANC streams as measured in the NSS Phase I
survey was 3 percent (4 percent) and 25 percent (22 percent), respectively. The calibrated percentage
(and percent length) of acidic and low ANC streams, therefore, were overestimated by factors of about
1.06 (1.01) and 1.13 (1.11), respectively.
276
-------
300
250
0-200
CD
""" 150
O
< 100
O
,_ 50
co
o>
> 0
-50
Mid-App Stream Reaches
ANC Comparison
-50 0 50 100 150 200 250 300
Phase 1 ANC (u.eq/L-1)
Mid-App Stream Reaches
[SO42-] Comparsion
0 50 100 150 200 250 300 350
Phase 1 [SO/-] (fieq L-1)
Mid-App Stream Reaches
([Ca2+]+[Mg2+]) Comparsion
— 400
0 50 100 150 200 250 300 350 400
Phase 1 ([Ca2+HMg2*]) (ueq L-I)
7.5
7.0
16.5
Q.
6.0
CO
-5.5
5.0
4.5
Mid-App Stream Reaches
pH Comparsion
4.5 5.0 5.5 6.0 6.5 7.0 7.5
Phase 1 pH
Figure 10-7. MAGIC calibrated values at year 0 and NSS Phase I values for ANC, sulfate,
calcium plus magnesium, and pH for the DDRP M-APP sample watersheds.
277
-------
1.0r
0.8
0.6
.1
O
a.
o
IX
o>
»3 0.4
ra
a
E
O
Mid-App Stream Reaches
ANC
0.0
•TOO
Phase 1
— Simulated Yr
0 100
ANC
200
L-I)
300
400
Mid-App Stream Reaches
1-0r
O 0.8
O
o.
o
0.6
0}
'••3 0.4
ro
3
E
O
o.o
Phase 1
Simulated Yr 0
"0 100 200 300 400 500 600
([Ca2+]+[Mg2*]) (jxeq !_"i)
Mid-App Stream Reaches
pH
to
O 0.8
O
S 0.6
yz 0.4
E
O °-2
0.0
JO
•^
o
a.
o
1.0 r
0.8
0.6
o.
0)
:P 0.4
—
rnase 1
— Simulation
Year 0
_
3
E
O
0.2
4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0
, pH
0.0
Mid-App Stream Reaches
[S042-]
Phase 1
— Simulated Yr 0
100 200 300
[SO*2'] (jieq L-i)
400
Figure 10-8. MAGIC calibrated values at year 0 and NSS Phase I values for ANC, calcium
plus magnesium, pH, and sulfate for the DDRP M-APP target population. Note
that this does not include all NSS-Phase I stream reaches in the M-APP.
278
-------
Table 10-6. Descriptive Statistics of Calibrated ANC, pH, Sulfate Concentration, and Calcium
Plus Magnesium Using MAGIC Compared with NSS Observed Values
Descriptive Statistics for MAGIC Calibrated Constituents
Variable
ANC
PH
SO4
Ca + Mg
Mean
77
6.50
172
242
St Dev
71
0.62
59
79
Min
-6
5.02
60
68
P25
20
6.19
120
178
Median
64
6.69
161
246
P75
111
7.27
210
325
Max
252
322
373
Descriptive Statistics for 1986 NSS Data (29 sites only, corresponding to MAGIC population)
Variable
ANC
pH
SO4
Ca + Mg
Mean
88
6.55
152
239
StDev
59
0.60
47
78
Min
-10
4.91
55
64
P25
48
6.41
110
179
Median
87
6.72
154
245
P75
139
6.96
178
322
Max
200
7.43
264
372
279
-------
Table 10.7. Population Estimates of Combined Number, Length and Percentages of Stream
Pleaches Having ANC Less Than Selected Reference Values at Downstream Nodes
During Spring Baseflow for NSS Phase I; All M-APP DDRP; DDRP Systems
Simulated (ANC Values Estimated from NSS-I); and DDRP Systems Simulated (ANC
Values Calculated from MAGIC Calibrations).
NSSa
DDRPb
DDRP Modelled
as Measured0
DDRP Modelled
as Calibrated6
Sample (n) 164
Target Population
(N) 25,715
Length (km) 69,569
ANC < 0 u.eq/L
(N) 326
(%)e 1.3.
Length (km) 2,324
(%)f 3.3
ANC < 50 |ieq/L
(N) 2,199
(%) 8.6
Length (km) 7,430
(%) 10.7
36
5,496
15,239
181
3.3
597
3.9
1,504
27.4
3,002
19.7
29
4,298
11,246
145
3.4
472
4.2
1,087
25.3
2,505
22.3
29
4,298
11,246
390
9.1
527
4.7
1,656
38.5
3,756
33.4
a National Stream Survey estimates for Subregions 1D, 2Cn, and 2Bn.
b DDRP sample and target population estimates for the Mid-Appalachian Region.
c Summary for NSS Phase I ANC values for stream reaches modelled by MAGIC.
d Summary for ANC values as calibrated by MAGIC for year 0.
9 Percentage of estimated target population number.
' Percentage of estimated target population length.
280
-------
10.6.4 Mid-Appalachian Projections
Projected changes in ANC, pH, sulfate, and calcium plus magnesium that might occur over a
100-year period under the three deposition scenarios are presented in this section.
10.6.4.1 Projected Values of ANC, pH, Sulfate, and Calcium Plus Magnesium
The three deposition scenarios are projected to have different effects on surface water
chemistry. These projected effects can be presented both as projections of absolute values of
constituents and as changes in concentrations (or values) of constituents. Although it is useful to
discuss both absolute values and changes, there are some biases of initial calibrated values of ANC
and sulfate concentration (Table 10-6); thus, for these constituents, the values that are likely to be
most meaningful are not the projected absolute values but rather, the projected changes. Projected
values represent flow-weighted annual average constituent concentrations.
Values of ANC are projected by MAGIC to decrease monotonically over 50 years under current
deposition levels (Scenario A), whereas they are projected to increase monotonically over 50 years
with a 50 percent reduction in current deposition (Scenario B) (Figure 10-9, Table 10-8). With a 50/20
percent change in deposition (Scenario C), ANC is projected to increase for the first 20 years, return to
approximately initial values (i.e., yr 0) after 50 years, and decrease after 100 years. The median
projected ANC values after 50 years are 36, 80, and 64 ^eq/L, respectively, for deposition Scenarios A,
B, and C (Table 10-8). The projected ANC values after 100 years for these three deposition scenarios
are 8, 77, and 45 /weq/L, respectively. The ANC is projected to decrease from yr 50 to yr 100 with all
three deposition scenarios (Table 10-8).
Values of pH are projected to show a pattern similar to that of ANC (Figure 10-10). The pH
values are projected to decrease monotonically over the 100-year period with current deposition rates
(Table 10-8). The median pH is projected to increase for the first 20 years with a 50 percent decrease
in deposition but then decrease over the next 80 years. Similar changes in pH are observed with the
50/20 percent deposition decrease, but the changes are greater over the latter 80-year period with this
deposition scenario. The median pH after 50 years with deposition scenarios A, B, and C are 6.4, 6.8,
and 6.7, respectively. After 100 years, the median pH values for these three deposition scenarios are
5.6, 6.8, and 6.5, respectively (Table 10-8). The pH changes are greater for all deposition scenarios at
the lower pH values (Figure 10-10). All streams with initial calibrated pH values less than about 6.5
exhibited greater changes in pH over time than those streams with initial calibrated pH over 6.5. The
cumulative distributions are similar over time regardless of deposition levels for streams with pH above
6.5. The lower quartile pH values after 50 years for current, 50 percent, and 50/20 percent deposition
are 5.6, 6.5, and 6.2, respectively; after 100 years, these values are 4.7, 6.4, and 5.5, respectively. At
the upper quartile, the changes after 50 and 100 years for all three deposition scenarios were to pH
values of 6.8, 7.0, 6.9, and 6.6, 7.0, and 6.8, respectively.
281
-------
1.0
O 0.8
*3
O
a.
2 0.6
a.
0>
I0'4
ZJ
O
0.0
Mid-App Stream Reaches
Deposition Scenario A
-100
100
200
ANC (fieq L-I)
300
400
1-Or
0.8
O
g-
0.6
CP
:p 0.4
JO
=>
E
O
Mid-App Stream Reaches
Deposition Scenario B
0.0'—
-100
Year 20
Year 50
0 100 200 300
ANC (|j,eq L-1)
400
I.Or
Mid-App Stream Reaches
Deposition Scenario C
0 100 200 300
ANC ([ieq L-1)
400
Figure 10-9. Projections of ANC for the M-APP stream target population under current sulfur
deposition (Scenario A), 50 percent deposition decrease (Scenario B), and 50/20
percent deposition decrease (Scenario C) over 100 years.
282
-------
Table 10-8. Descriptive Statistics of Projected ANC, pH, Sulfate, Percent Sulfur Retention,
and Calcium and Magnesium Using MAGIC for Current and Decreased
Deposition
CURRENT DEPOSITION (Scenario A)
Model
ANC
YrO
Yr20
Yr50
Yr 100
YrO
Yr20
Yr50
Yr 100
SO4
YrO
Yr20
Yr50
Yr 100
Percent S
Yr20
Yr50
Yr 100
Ca + Mg
YrO
Yr20
Yr50
Yr 100
MEAN
77
66
51
31
6.50
6.36
6.14
5.69
172
208
232
239
Retention
14
4
1
242
266
273
244
STD DEV
71
69
64
60
0.62
0.71
0.81
0.95
59
66
67
70
12
7
7
79
87
94
96
MIN
-6
-9
-15
-21
5.02
4.92
4.69
4.49
60
100
128
128
-6
-13
-15
68
99
114
100
P25
20
12
4
-15
6.19
6.01
5.65
4.69
120
148
191
192
6
0.7
-0.9
178
207
183
166
MED
64
49
36
8
6.69
6.57
6.42
5.65
161
194
218
226
12
4
2
246
254
267
238
P75
111
102
83
49
6.92
6.90
6.80
6.57
210
236
267
276
24
11
9
325
345
345
298
MAX
252
238
216
183
7.27
7.25
7.20
7.13
322
381
389
389
48
21
12
373
414
467
472
(Continued)
283
-------
Table 10-8. Continued
50% DECREASE (Scenario B)
Model
ANC
YrO
Yr20
YrSO
YrlOO
EH
YrO
Yr20
YrSO
YrlOO
SO4
YrO
Yr20
YrSO
YrlOO
Percent S
Yr20
YrSO
YrlOO
Ca + Ma
YrO
Yr20
YrSO
YrlOO
MEAN STD DEV
77
97
93
85
6.50
6.72
6.70
6.63
172
146
130
125
Retention
-20
-8
-3
242
240
221
209
71
71
69
67
0.62
0.42
0.43
0.49
59
45
35
34
14
12
11
79
81
80
78
MIN
-6
5
2
0
5.02
5.71
5.60
5.43
60
80
69
67
-54
-39
-34
68
93
98
91
P25
20
45
40
30
6.19
6.54
6.50
6.37
120
118
109
104
-26
-15
-8
178
170
151
142
MED
64
84
80
77
6.69
6.81
6.79
6.77
161
134
123
120
-16
-3
0.1
246
237
219
210
P75
111
132
129
122
6.92
7.01
7.00
6.98
210
168
143
141
-13
0.3
4
325
310
253
235
MAX
252
260
249
234
7.27
7.29
7.26
7.24
322
255
202
195
15
12
11
373
388
391
376
(Continued)
284
-------
Table 10-8. Continued
50%/20% DECREASE (Scenario C)
Model
ANC
YrO
Yr20
Yr50
Yr 100
EH
YrO
Yr20
Yr50
Yr 100
504
YrO
Yr20
Yr50
Yr 100
Percent S
Yr20
Yr50
Yr 100
Ca + Mq
YrO
Yr20
Yr50
Yr 100
MEAN
77
95
72
55
6.50
6.71
6.46
6.19
172
149
172
191
Retention
-13
11
1
242
240
239
240
STDDEV
71
70
67
64
0.62
0.43
0.63
0.81
59
45
47
54
13
9
8
79
81
83
87
MIN
-6
4
-5
-13
5.02
5.67
5.08
4.76
60
83
100
104
-45
-12
-21
68
93
109
102
P25
20
43
19
0.8
6.19
6.53
6.19
5.48
120
120
142
154
-19
8
-2
178
171
164
158
MED
64
82
64
45
6.69
6.80
6.69
6.53
161
138
160
180
-8
12
4
246
238
233
238
P75
111
131
101
78
6.92
7.01
6.89
6.78
210
169
200
219
-5
17
8
325
311
297
298
MAX
252
258
234
209
7.27
7.28
7.24
7.19
322
261
283
311
21
32
13
373
389
412
431
285
-------
1.0
O 0.8
o
0.
o
0.6
03
I0'4
r>
O
0.0
Mid-App Stream Reaches
Deposition Scenario A
4.S 5.0 5.5 6.0 6.5 7.0 7.5 8.0
PH
Mid-App Stream Reaches
Deposition Scenario B
4.5 5.0 5.5 6.0 6.5 7.0 7.5
8.0
Mid-App Stream Reaches
Deposition Scenario C
4.5 5.0 5.5 6.0 6.5 7.0 7.5 8.0
Figure 10-10. Projections of pH for the M-APP stream target population under current sulfur
deposition (Scenario A), 50 percent deposition decrease (Scenario B), and
50/20 percent deposition decrease (Scenario C) over 100 years.
286
-------
Sulfate concentrations are projected to increase over time under current deposition, decrease
under the 50 percent deposition scenario, and decrease for the first 20 years and then increase for the
next 80 years with the 50/20 percent deposition scenario (Figure 10-11).. The median percent sulfur
retention in the watersheds is projected to approach steady state after about 100 years under both
deposition Scenarios A and B (Table 10-8). Under Scenario A, however, the median sulfate concen-
trations are projected to continue to increase in the streams as they approach sulfur steady state,
whereas with Scenario B, the watersheds are projected to desorb sulfate as they approach a new
sulfur steady state and the stream sulfate concentrations decrease (Table 10-8). With Scenario C, the
change in the median percent sulfur retention indicates the watersheds initially desorb sulfate, with
median stream sulfate concentration decreasing over the first 20 years; then the median sulfate
concentrations in the streams are projected to increase as the watersheds approach a new sulfur
steady state (Table 10-8).
The model projects calcium plus magnesium concentrations to increase for the first 50 years
under deposition Scenario A and then decrease for the next 50 years (Figure 10-12). Under Scenario
B, calcium plus magnesium concentrations are projected to decrease monotonically over the 100-year
period, whereas for Scenario C, calcium plus magnesium concentrations are projected to decrease
slightly for the first 50 years and then increase for the next 50 years (Table 10-8). The changes
projected for Scenario C are relatively small.
The results of the model projections can also be represented as (1) the increase or decrease in
the number of stream reaches with ANC < 0 ^weq/L or < 50 fieq/L at the lower reach nodes (Table
10-9), (2) the percentage of the DDRP target population with ANC < 0 ^eq/L or < 50 ^eq/L at the
lower reach nodes (Table 10-10), or (3) the reach length in the DDRP target population with ANC < 0
yweq/L or < 50 /^eq/L based on projections at the lower reach nodes (Table 10-11). Statistical tests of
significance (see Section 6) can be performed to determine if the changes at some target year (e.g.,
year 50) are significantly different from zero. Using the procedures presented in Section 6, we have
performed such tests for the change in the number of systems with ANC < 0 ^weq/L and ANC < 50
neaJL at the lower reach nodes. In Section 10.8.5.2, we discuss the important implications of making
projections of reach chemistry based only on the chemistry at the lower nodes of each reach.
For deposition Scenario B, MAGIC projects a decrease in the number of acidic systems and the
number of systems with ANC < 50 ^weq/L (in some cases to zero acidic systems), but with p > .16 for
all cases (Table 10-9), where the null hypothesis is that the projected change is not significantly
different from zero. For Scenario C, MAGIC projects first a decrease in the number of acidic and low
ANC systems (p = 16 at 20 years), then an increase in the number of such systems (.07 ^ p ^ .09 at
50 and 100 years). For the current deposition scenario (Scenario A), all cases of projected changes in
the number of acidic systems at 20, 50, and 100 years have p < .05 (Table 10-9). For the systems
with ANC < 50 /*eq/L, p = .01 for the projected changes at 50 and 100 years.
The projections at current deposition of acidic systems at 50 and 100 years correspond to 11
percent and 24 percent, respectively, in the number of stream reaches, based on lower node
chemistry (Table 10-10) and to 1,298 km (at 50 years) and 1,886 km of stream reach length (at 100
years) (Table 10-11). The projections for stream reaches with lower node ANC < 50^eq/L at 50 and
287
-------
1.0r
Mid-App Stream Reaches
Deposition Scenario A
100
200 300
[SO,2-] (jieq L-1)
400
0.0
Mid-App Stream Reaches
Deposition Scenario B
400
1.0r
Mid-App Stream Reaches
Deposition Scenario C
400
Ftgure 10-11. Projections of sulfate concentration for the M-APP stream target population
under current sulfur deposition (Scenario A), 50 percent deposition decrease
(Scenario B), and 50/20 percent deposition decrease (Scenario C) over 100
years.
288
-------
Mid-App Stream Reaches
Deposition Scenario A
0 100 200 300 400 500 600
([Ca2+]+[Mg2D (neq 1_-1)
Mid-App Stream Reaches
Deposition Scenario B
100 200 300 400 500
([Ca2+]+[Mg2*]) (jieq L"0
600
1.0
O 0.8
O
&0.6
Q.
O
V D.4
i
O 0-2
0.0
Mid-App Stream Reaches
Deposition Scenario C
100 200 300 400 500
([Ca2+]+tMg2*]) (jieq L-1)
600
Figure 10-12. Projections of calcium plus magnesium for the M-APP stream target population
under current sulfur deposition (Scenario A), 50 percent deposition decrease
(Scenario B), and 50/20 percent deposition decrease (Scenario C) over 100
years.
289
-------
Table 10-9. Change Projected by the MAGIC Model In the Number of M-APP DDRP
Lower Reach Nodes Having (A) ANC < 0 or (B) ANC < 50 fiec\/L
ANC < 0 tiey/L
Depos'rtion Scenario8
Year
20
50
100
Current
Scenario A
50
p=.02
347
p=.04
906
p=.02
50% Decrease
Scenario B
-145
p=.16
-145
p=.16
-130
p=.19
50%/20% Decrease
Scenario C
-145
p=.16
30
p=.07
192
p=.09
B
ANC < 50 uea/L
Year
Current
Scenario A
Deposition Scenario8
50% Decrease
Scenario B
50%/20% Decrease
Scenario C
20
50
100
229
p=.16
1240
p=.01
1300
p=.01
-145
p=.16
-145
p=.16
83
NS
-145
p=.16
229
p=.07
546
p=.08
The number given is the estimated difference in number of population lower reach nodes between the listed year
and year 0 for tha same MAGIC model runs. Each ANC value is calculated as the NSS observed value plus the change
in model ANC since year zero.
in each case, the difference is compared against a statistical estimate of the standard deviation of the difference,
and a p-value is given underneath the difference. The test is a one-sided hypothesis test where the null hypothesis
Is that the difference is not significantly different from zero. The estimate of the variability is designed to
capture the variability of the inputs into the model, the sensitivity of the model to the input variability, the
variability of the model between scenarios, and the variability involved in extrapolating to the population estimates.
290
-------
Table 10-10. Percentage of M-APP Target Population Stream Reaches Projected by the MAGIC
Model to Have (A) ANC < 0 or (B) ANC < 50 /*eq/L at the Lower Reach Nodes
A
ANC < 0 weq/L
Year Current
Scenario A
Oa 3
20b 5
50b 11
100b 24
Deposition Scenario
50% Decrease
Scenario B
3
0
0
0
50%/20% Decrease
Scenario C
3
0
4
8
B
ANC < 50 weq/L
Year Current
Scenario A
Oa 25
20b 31
50b 54
100b 56
Deposition Scenario
50% Decrease
Scenario B
25
22
22
27
50%/20% Decrease
Scenario C
25
22
31
38
a Estimates from National Stream Survey for the target population stream reaches corresponding to the watersheds modelled
(see Table 10-7).
b Estimates at 20, 50 and 100 years are computed by adding change in ANC projected by MAGIC to ANC observed by
Kaufmann et al. (1988) for the National Stream Survey.
291
-------
Table 10-11. Length (km) of M-APP Target Population Stream Reaches Projected by the MAGIC
Model to Have (A) ANC < 0 or (B) ANC < 50 fieq/L at the Lower Reach Nodes
A
ANC < 0 /tea/L
Year
Oa
20b
50b
100b
Current
Scenario A
472
773
1298
1886
Deposition Scenario
50% Decrease 50%/20% Decrease
Scenario B Scenario C
472 472
0 0
0 645
80 1016 -
B
ANC < 50 «ea/L
Year
Oa
20b
50b
100b
Current
Scenario A
2505
2777
5831
6178
Deposition Scenario
50% Decrease 50%/20% Decrease
Scenario B Scenario C
2505 - 2505
2221 2221
2221 2777
2493 4694
a Estimates from the National Stream Survey for the target population stream lengths corresponding to the watersheds
modelled (see Table 10-7).
b Estimates at 20, 50. and 100 years are computed by adding change in ANC projected by MAGIC to ANC observed by
Kaufmann et al. (1988) for the National Stream Survey.
292
-------
100 years are over 50 percent of the number of stream reaches (54 percent and 56 percent, respec-
tively). These projections are equivalent to 5,831 km and 6j178 km of stream reaches, respectively, at
50 and 100 years (Table 10-11). Over half the DDRP target population of stream reaches (at the lower
nodes, see Section 10.8.5.2) are projected to have ANC < 50 ^eq/L and thus, presumably, to be
susceptible to episodic acidification to ANC < O^eq/L (Eshleman, 1988).
10.6.4.2 Projected Rates of Change
The projected changes in ANC and pH values over the next 100 years can be displayed as box
and whisker plots (Figures 10-13 through 10-18), which illustrate how both the target population
constituent medians and interquartile ranges vary through time.
The change in median ANC concentration projected for current deposition (Scenario A) was -28
,weq/L (64 ^eq/L to 36 ^ueq/L) after 50 years and -56 ^eq/L after 100 years (Table 10-8). The projected
rate of ANC change was relatively constant over time (Figure 10-13). For the 50 percent decreased
deposition scenario (Scenario B), the projected change in median ANC was +16 ^weq/L (64 to 80
^weq/L) at year 50, and +13lweq/L (64 to 77/^eq/L) at year 100 (Figure 10-14). With the 50/20 percent
deposition scenario (Scenario C), the change in median ANC was 0 ^ueq/L or equivalent to the initial
calibrated median ANC at 50 years and -19^eq/L (64 to 45,aeq/L) at 100 years (Figure 10-15). The
projected change in median ANC varied among ANC groups (Table 10-12). For deposition Scenario
A, the projected changes in median ANC for streams with an initial ANC < 25 /*eq/L were about -7
fieq/L and -16 fieq/L after 50 and 100 years, respectively. With an initial ANC concentration of 25 to
100 fteq/L, the projected changes were -33 and -61 /ieq/L, respectively. The projected change in
ANC was intermediate for streams with an initial calibrated ANC between 100 and 40O/*eq/L (Table
10-12). These changes are shown graphically in Figure 10-19. For Scenario B, the greatest projected
ANC increase occurred in streams with an initial calibrated ANC between 100 and 400^eq/L (Figure
10-19). For Scenario C, almost no change was projected for streams with an initial calibrated ANC
< 25 jMeq/L, whereas for streams with an initial calibrated ANC between 25 and 100 ,weq/L the
projected change was -7 fieq/L (Table 10-12). After 100 years, the greatest ANC change (-27yweq/L)
was projected for streams with initial calibrated ANC between 100 and 400
For deposition Scenario A, the projected change in median pH was -0.25 units at 50 years and
-1.0 units during the next 100-years (Table 10-8, Figure 10-16). Median pH for Scenario B remains
nearly constant over the 100-year period (Table 10-8, Figure 10-17), increasing by 0.10 units between
years 0 and 50, but then declining by .02 units between years 50 and 100. For Scenario C, median
pH is projected to change little during the initial 50 years of simulation, but to decline by about 0.20
units between years 50 and 100 (Table 10-8, Figure 10-18).
The change in median ANC is a population attribute, not an individual watershed attribute. The
change in ANC for each individual watershed was relatively independent of the initial, calibrated ANC
values. When watershed groups (e.g., < 25, 25-100, 1 00-200 fieq/L ANC) are considered, however,
different watershed characteristics and differential population weights result in different changes in
median ANC, not as a function of the initial ANC, but rather as a function of the initial ANC group to
which the watershed was assigned.
293
-------
ANC vs. TIME
Constant Deposition
Model * MAGIC
-
o
CM
a:
o
LO
Mid-
Appalachian
Region
O
o
>-
Figure 10-13. Box and whisker plots of projected ANC for M-APP streams under current
deposition (Scenario A).
294
-------
ANC vs. TIME
SOX Decrease
Model - MAGIC
D"
Q)
D
O
250
200
150
100
50
0
-50
Upper
Q3
Mean
Median
Qi
Lower
o
>-
O
CM
O
m
a:
Mid-
Appalachian
Region
o
o
Figure 10-14. Box and whisker plots of projected ANC for M-APP streams under a 50 percent
deposition decrease (Scenario B).
295
-------
ANC vs. TIME
50X/20X Decrease
Model - MAGIC
cr
•
O
CN
O
m
Mid-'
Appalachian
Region
O
O
Figure 10-15. Box and whisker plots of projected ANC for M-APP streams under a 50/20
percent deposition decrease (Scenario C).
296
-------
pH vs. TIME
Constant Deposition
Model - MAGIC
8r
7
a_
4-
Upper
Q3
Mean
Median
Qi
Lower
o
CM
O
in
Mid-
Appalachian
Region
o
o
Figure 10-16. Box and whisker plots of projected pH for M-APP streams under current
deposition (Scenario A).
297
-------
pH vs. TIME
SOX Decrease
Model - MAGIC
Upper
Mean
Median
Qi
Lower
Mid-
Appalachian
Region
8
7
o
CM
o
LO
O
o
a:
Figure 10-17. Box and whisker plots of projected pH for M-APP streams under a 50 percent
deposition decrease (Scenario B).
298
-------
pH vs. TIME
50X/20X Decrease
Model - MAGIC
Mid-
Appalachian
Region
8
7
i.
4
Upper
Q3
Mean
Median
Qi
Lower
>-
o
CM
Of
o
lO
O
o
Figure 10-18. Box and whisker plots of projected phi for M-APP streams under a 50/20
percent deposition decrease (Scenario C).
299
-------
Table 10-12. Changes in Median ANC, pl-i, Sulfate Concentration, and Calcium plus Magnesium
Concentration over 50 and 100-Year Periods as a Function of the Initial NSS ANC
Groups and Deposition Scenarios
NSS ANC Group
Mode! Output
Variable
< 25
25 - 100
100 - 400
Scenario
ANC (Meq/L)
50 yr
100 yr
EhL
50 yr
100 yr
SO4 (Meq/L)
50 yr
100 yr
Aa
-7
-16
-0.4
-0.7
28
31
Bb
15
10
0.7
0.5
-39
-41
cc
0.5
-7
0.1
-0.4
-8
2
A
-33
-61
-0.3
-0.9
35
43
Scenario
B
9
6
0.1
0.0
-6
-63
C
-7
-26
-0.1
-0.2
-23
-2
Scenario
A
-23
^6
-0.1
-0.3
106
116
B
24
17
0.1
0.1
-27
-34
C
-5
-27
-0.0
-0.1
17
56
(Ca + Ma) (uea/L)
50 yr
100 yr
-4
-30
* Deposition Scenario A =
Deposition Scenario B =
c Deposition Scenario C =
-32
-37
-22
-24
14
5
-22
-32
-10
-6
28
-37
-82
-96
-38
-37
current deposition.
50 percent decrease.
50/20 percent decrease.
300
-------
Change
in Median ANC
Year 0 to Year 50
Model - MAGIC
Mid-
Appalachian
Region rv
w
Constant 50X 50//20X
Deposition Decrease Decrease
\ 30 r
cr
CD 9O
^ Z.U
~ 10-
CD
f~~D f~\
C U &g$&::::$.:&:
5 -10 -
O >>:::^:iS
, -20-
O svvSX;
< ' ' ' '•'
30r 30
9 0 ::;::;;;:-':;;; ? 0
z.vj n J-.-.--.-.-.-.-J ^.vj
-If) i^>%; :W:W: 1Q
H :-:-::-:-::-:-;: •:••:•:•-.-••.-:-- :•:••:•:-•:•:•-: pi
U U
>^ -10 - -10
li -20 - -20
-10 - -"50
r-
_
$-:;::::$£M
-
-
moo LOOO LOOO
CMOO CNOO ^|Op
x/ i
OJ '
•NT v , ^sT " i ^J
§ S g Sg
ANC Group ANC Group ANC Group
(ueq/L) (ueq/L) (ueq/L)
Figure 10-19. Projected change in median ANC concentration over 50 years for all three
deposition scenarios as a function of the NSS Phase i ANC used to classify
DDRP streams in the target population.
301
-------
Projected changes in pH are closely related to year 0 pH and ANC of the streams. The greatest
changes in pH are projected to occur in streams with low ANC (Figure 10-20), although as shown by
Figure 10-21, the projected decreases in streams with an initial pH between 6.0 and about 6.75 are
greater than for streams with initial pH < 6.0. This pattern can be attributed to two factors: (1) as pH
decreases below 6.0, further decreases in pH are partially buffered by dissolution of aluminum, and
(2) pH is a logarithmic, rather than a linear scale. Thus, although increases in H+ are greatest in
those surface waters with the lowest pH, corresponding changes in pH are smaller. Under current
deposition, pH is projected to decrease by 0.1 to 1.9 units by year 100. For the 50 percent decreased
deposition scenario, pH is projected to increase in most systems that currently have pH > 6.5;
streams that presently have a high pH are projected to change very little over the next 100 years
(Figure 10-21). For the 50/20 percent scenario, pH is projected to change very little between years 0
and 50, but to decline in most systems between years 50 and 100; the largest declines are projected
to occur in those streams with year 0 pH between 6.0 and 6.5 (Figure 10-21).
Projected changes in sulfate reflect the continuation of an ongoing trend of soils and runoff to
come to steady state with atmospheric sulfur deposition. Currently some soils in DDRP watersheds in
the M-APP Region are retaining a significant portion of incident sulfur by adsorption on soils; as soils
continue to adsorb sulfate and approach a dynamic equilibrium with current levels of deposition,
sulfate concentrations in drainage waters will increase to steady-state.concentrations. Conversely, if
deposition is reduced, soils will desorb sulfate and come to steady state with the lower level of inputs.
For the M-APP scenario of continued deposition at current loadings (Scenario A), the projected
change in sulfate during the next 50 years is +57/*eq/L (from 161 to 218 /*eq/L) and +65 ^eq/L
during the 100-year simulation period (Table 10-8). For the 50 percent deposition scenario (Scenario
B), median sulfate concentration is projected to decrease by SS^eq/L during the next 50 years and by
41 /
-------
Change in Median pH
Year 0 to Y
Model- MAC
Constant
Deposition
0.8r 0.8
0^™* ^\ /*^
.6 - 0.6
c? 0.4 - 0.4
g 0.2- 0.2
-^ &£#ffi^W3 °
^-0.2 Ililil -0.2
OA ££i££ n A
. "r U . "f
LO 0 0
Csl O O
v s I
ANC Group
(ueq/L)
ear 50
3IC
SOX
Decrease
r 0.8
-TxX'*-v
$m
_
0.6
0.4
0.2
h 1 Tt 0
-0.2
n /
V^f • 1^
Mid-
Appalachian
Regionrx
50X/20X
Decrease
-
in o o in o o
CM O O (N O O
"^ 0 ^0
OJ (-) CN] (-)
00 0 °-
ANC Group ANC Group
(ueq/L) (ueq/L)
Figure 10-20. Change in median pH concentration over 50 years for all three deposition
scenarios as a function of the NSS Phase I ANC used to classify DDRP streams
in the target populations.
303
-------
Mid-App Stream Reaches
Deposition Scenario A
Year = 50
1.0
0.5
o.o
••0.5
-to
-15
-2.0
8>
4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5
Year 0 Model pH
Mid-App Stream Reaches
Deposition Scenario A
Year = 100
tOr
0.5
0.0
CL
-0.5
-1.0
-1.5
-2.0
4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5
Year 0 Model pH
i.
to
0.5
0.0
-0.5
•to
•ts
-2.0
Deposition Scenario B
Year = 50
O 00
4.0 4.5 5,0 5.5 6.0 6.5 7.0 7.5
Year 0 Model pH
to
0.5
0.0
L-0.5
I
-1.0
-1.5
Deposition Scenario B
Year - 100
* °D°
o 8 e>
4.0 4.5 5.0 5.5 8.0 6.5 7.0 7.5
Year 0 Model pH
to
0.5
0.0
L-0.5J
I
•to
•ts
-2.0
Deposition Scenario C
Year - 50
4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5
Year 0 Model pH
tOr
0.5
0.0
•"0.5
-1.0
-1.5
-2.0
Deposition Scenario C
Year - 100
4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5
Year 0 Model pH
Figure 10-21. Change in pH for sampled streams in the DDRP M-APP Region as a function of
the initial calibrated pH at year 0.
304
-------
10.6.4.3 Comparison With ILWAS Projections for Two Watersheds
Staff of Tetra-Tech, Inc., used the Integrated Lake-Watershed Acidification (ILWAS) Model to
perform projections of potential future effects of sulfur deposition for two of the DDRP M-APP
watersheds (R. Munson, pers. comm.). They performed simulations for the DDRP watersheds of
Johnson Run, West Virginia {ID 2C046033; observed ANC = 6 jieq/L) and North Branch Rock Run,
Pennsylvania (ID 1D029023; observed ANC = 32fieq/L). They used data for surface water chemistry,
watershed characteristics, soil attributes and chemistry, annual runoff, and current atmospheric
deposition chemistry and volume provided by DDRP staff. They performed simulations using three
deposition scenarios provided by the Electric Power Research Institute (EPRI) (R. Munson, pers.
comm.). Of the three scenarios, only one corresponded directly to a scenario used by DDRP. This
was the deposition scenario of continuing deposition at current levels (Scenario A). Tetra-Tech, Inc.,
performed projections for periods of up to 50 years.
Here, we briefly examine comparisons between the ILWAS and MAGIC projections for the two
watersheds using flow-weighted average annual values for ANC, the primary variable of interest in the
DDRP. For ANC, it is best to compare projected changes (rather than absolute values) because
complications may arise from calibration offsets for current values. As noted earlier, in our reporting of
final projected ANC values we combine projected changes with current values as observed by the
National Stream Survey (Kaufmann et al., 1988). This comparison of model projections parallels that
approach.
For Johnson Run, West Virginia (ID 2C046033), both MAGIC and ILWAS projected very slight
acidification (loss of ANC) in the future under current deposition. Projections with ILWAS indicated
decreases in ANC of 1 and 2 neq/L at 20 and 50 years, respectively. Projections with MAGIC
indicated ANC decreases of 2 and 5 [j.eq/L at 20 and 50 years, respectively.
For North Branch Rock Run, Pennsylvania (ID 1D029023), projections from MAGIC and ILWAS
both indicated acidification slightly greater than that for Johnson Run under current deposition
loadings. At 20 years, MAGIC projected a decrease of 8 neq/L, and ILWAS a decrease of 6 neq/L. At
50 years, MAGIC and ILWAS projected decreases of 15 and 11 neq/L, respectively.
Two primary observations can be made from these comparisons. First, both models project
future acidification for the watersheds (very slight for one, more pronounced for the other) under
current loads of atmospheric deposition. Second, the changes in ANC projected by the models for
the two watersheds are in close correspondence, with differences in projections ranging from 1 to 4
|ieq/L
10.7 DISCUSSION
10.7.1 Projections of Future Changes in Acid-Base Surface Water Chemistry
One of the objectives of the DDRP Level III analyses is to estimate the general direction and
relative magnitude of possible future changes in surface water acid-base chemistry as a function of
305
-------
alternative sulfur deposition scenarios. Making relative comparisons among different management or
control alternatives represents one of the most effective uses of models. These comparisons are
based on relative changes in the population attributes of M-APP stream reaches as a function of the
three deposition scenarios. Annual typical year meteorological and deposition data were repeated for
100 years to provide a consistent base for assessing the relative magnitude and direction of the
change in stream chemistry of M-APP watersheds and to permit relative comparisons among deposi-
tion scenarios. These projections are not and were not intended to be forecasts of future conditions.
Forecasts relate to the actual probability that an event or condition will occur. The relative magnitude
and direction of change can be placed in perspective, however, by comparing the projections with
other information on M-APP watersheds, such as monitoring and survey data.
Dynamic watershed models represent an integration of our knowledge and working hypotheses
of how watershed processes control changes in surface water acid-base chemistry (Jenne et al., 1989;
Thornton et al., 1990). The observed differences in process inclusion/representation among models
reflect the different philosophies among model developers regarding both the quantitative importance
of individual processes and the different descriptions of the nature of those processes. A discussion
and evaluation of these alternative hypotheses and underlying assumptions, as well as other uncertain-
ties, provides information on our level of understanding and assists in identifying important remaining
questions. Tabfe 10-13 lists the implications of a variety of factors on estimates of future change in
ANC and pH. Section 10.8 discusses some of these factors and their implications on conclusions that
can be drawn from comparisons among model projections of different deposition scenarios.
10.7.2 Systems of Interest
The Mid-Atlantic highlands consists of the Mid-Atlantic parts of the Blue Ridge, Ridge and Valley,
and Appalachian Plateaus, an area bounded by the Virginia/North Carolina border on the south and
the Adirondack Mountains of northern New York on the north (Kaufmann et al., 1988). This area
includes the DDRP Mid-Appalachian Region. In the Mid-Atlantic highlands, Kaufmann et al. (1988)
found acidic streams (excluding streams acidic due to acid mine drainage, which were removed from
this analysis) only in forested watersheds with areas less than 30 km2. These forested watersheds
typically occur at relatively high elevations within individual physiographic regions, and are located on
ridges in the Ridge and Valley Province and in upland areas of the Appalachian Plateau. Kaufmann et
al. (1988) concluded that the most probable source of acidity in the acidic streams was acidic
deposition. In the acidic systems, sulfate.was the dominant anion and stream sulfate concentrations
could be accounted for based on evapoconcentration of sulfate deposition, making substantial
watershed contributions unlikely. Nitrate and organic anion concentrations were very low and
contributed little to chronic surface water acidity (Kaufmann et al., 1988).
The major factor controlling surface water ANC concentrations in the Mid-Atlantic highlands,
based on current chemistry, is apparently neutralization of acidic inputs through mineral weathering or
cation exchange with the concomitant release of base cations (predominantly calcium) (Kaufmann et
al., 1988). The acidic streams had mean sulfate concentrations of about 150 ^eq/L and mean base
cation concentrations of about 135^eq/L.
306
-------
Table 10-13. Effects of Critical Assumptions and Uncertainties on Projected Rates of Change
Assumptions Resulting in Under-
Estimates of ANC and pH Changes
Assumptions Resulting in Over-
Estimates of ANC and pH Changes
1. Mineral weathering overestimated 1.
2. Nitrate assimilation overestimated 2.
3. Total sulfur deposition underestimated 3.
4. Calibrated ANC greater than observed 4.
5. Watershed land use changes 5.
6. Desorption is not the reverse of 6.
adsorption-sulfate is irreversibly
bound
7. Biotic uptake/assimilation reducing 7.
available base cation pool
8. Effects of distribution extremes over-
smoothed through aggregation
Mineral weathering underestimated
Organic acids buffer surface water
chemistry
Total sulfur deposition overestimated
Calibrated ANC less than observed
Watershed land use changes
Desorption is not the reverse of
adsorption-hysteresis-related delays in
change
Weathering and sulfate adsorption
increased by decreased soil pH
307
-------
The DDRP M-APP target population simulated by MAGIC "represents streams with measured -
ANC < 200 /teq/L, and watersheds having areas < 30 km2 and showing no evidence of acid mine
drainage or other significant internal sources of sulfur. The DDRP stream systems are predominantly
higher elevation, forested watersheds. The initial calibrated median ANC for the stream target
population was 64/^eq/L, with median sulfate and base cation concentrations of 161 and 246 ^eq/L,
respectively. The lower quartile and median pH values were 6.2 and 6.7, respectively. This target
population of watersheds represents the streams that are likely to respond to acidic deposition inputs
over the next several decades.
10.7.3 Relative.' Magnitude and Direction of Change
Under current levels of deposition, the direction of change projected by model simulations was a
decrease in ANC. Simulations result in a statistically significant increase in the percent of acidic
systems at 20, 50, and 100 years. With decreased deposition, there was a general projected trend of
increasing ANC, and a decrease in the percentage of currently acidic systems. For the 50/20 percent
deposition scenario, projected changes in sulfate paralleled changes in sulfur deposition (i.e., a
decrease, then a substantial increase in sulfate concentration), and resulted in changes in pH and
ANC that were inversely related to the changes in sulfate. The general direction of the change in
surface water acid-base chemistry projected by MAGIC for DDRP M-APP watersheds, in relation to
changes in sulfur deposition inputs, is consistent with current theories of surface water acidification,
and with laboratory and field observations (Church and Turner, 1986; Baker et al., 1991; Galloway et
al., 1983; NAS, 1984; Schindler, 1986, 1988; Turner et al., 1990; Wright et al., 1990).
The rates and magnitudes of chemical changes projected by MAGIC for surface waters in the
M-APP Region are, in general, consistent with those observed at monitoring sites within the region.
For the current deposition scenario, median projected changes for ANC and sulfate between years 0
and 50 are -0.56 /*eq/L/yr and +1.1 /teq/L/yr, respectively. In comparison, for two stream systems
developed over sedimentary bedrock in the Blue Ridge Province of Virginia, that have soil and other
watershed characteristics comparable to those of the modeled DDRP target systems, sulfate concen-
trations have increased about 2 /*eq/L7yr and ANC has decreased by 0.5 to 0.7 ^ueq/Uyr over the past
10 years (Ryan et al., 1989). Sulfate increases of similar magnitude have been reported for Coweeta
streams; sulfate concentrations have increased and ANC has decreased to the point that sulfate is
now the dominant anion in some of these streams (Waide and Swank, 1987). Analysis of data for
USGS Hydrologic Bench Mark stations in the Mid-Atlantic region indicates that sulfate increases were
statistically significant in low ANC streams in the region (Smith and Alexander, 1983). There was no
statistically significant trend in sulfate concentration, however, for higher ANC USGS stream systems in
the southern Mid-Atlantic highlands. The rates and magnitude of change projected by MAGIC are
within the range of measured rates observed under current deposition rates. Because the DDRP
target population is limited to sites with ANC < 200 /
-------
the surveyed streams were acidic (ANC < 0 ^eq/L). Watersheds in these regions of Virginia are, on
average, currently retaining sulfur (about 68 percent net annual retention), with an estimated steady-
state suifate concentration of 220 fieq/L based on current deposition levels for the region. The median
net annual sulfur, retention for Mid-Appalachian watersheds, based on the DDRP analyses, is currently
40 percent but is projected to be near steady state in 50 years (Sections 7 and 9 of this report).
Webb et al. (1989) used empirical steady-state models to investigate future changes in the
Virginia brook trout streams. Assuming F factors of 0.4 and 0.8, they projected median ANC values for
sulfur steady state of -38 and 22 ^weq/L, respectively. The percentage of acidic systems for these two
F factors (0.4 and 0.8) were 88 percent and 32 percent acidic streams, respectively, when stream
suifate concentrations reach steady state (Webb et al., 1989). Projected median ANC values after 50
and 100 years, using MAGIC for current deposition, were 36 and 8 peq/L, respectively. The percen-
tages of acidic systems after 50 and 100 years were 11 percent and 24 percent (Table 10-9), or an
increase of 8 percent and 21 percent, respectively. The projected percentages of low ANC streams
(ANC < 50 ,aeq/l_) after 50 and 100 years were 54 percent and 56 percent, or an increase of 29 per-
cent and 31 percent, respectively. The projected, relative changes are consistent among these two
approaches.
10.7.4 Relationships Among Model Projections and Watershed Processes
10.7.4.1 Sulfur Retention Processes
10.7.4.1.1 Adsorption-
The MAGIC subroutine for suifate retention is based on two assumptions; the first is that suifate
adsorption is the only terrestrial process influencing watershed sulfur retention/release. The second
assumption, which affects projections of .watershed response when deposition is reduced, is that
adsorption is completely reversible. The review of sulfur retention processes in soils included with
DDRP analyses for the NE and SBRP (Church et al., 1989, Section 3) noted that a variety of processes
can retain or release suifate in terrestrial systems. That review concluded, however, that adsorption
was the principal net retention mechanism in most upland terrestrial systems. The adsorption capacity
of soils varies widely on both local and regional scales, and is related to soil pH, Fe, and Al content,
texture, and organic matter. Similarly, the reported reversibility of adsorption varies widely; suifate is
readily and completely desorbed from some soils (Weaver et al., 1985; Sanders and Tinker, 1975),
whereas less than 10 percent of suifate can be displaced from some highly weathered tropical soils
(Bornemisza and Llanos, 1967). Reversibility of sorption has not been well characterized for soils in
the M-APP Region and contributes to uncertainty in model projections; to the extent that suifate is
irreversibly sorbed to soils, recovery time in response to reductions in sulfur deposition will be
overestimated.
10.7.4.1.2 Retention of organic sulfur —
Along with adsorption, sulfur can be retained in soils by incorporation into soil organic matter. If
there is significant net retention or release of sulfur from organics, model estimates will over- or under-
309
-------
estimate actual watershed response time.. Any assessment of apparent reversibility of sorption and/or
hysteresis in desorption can similarly be confounded by changes in the rates of sulfur immobilization
and/or mineralization in soil organic matter. Although soil sulfur occurs predominantly in organic
forms in most soils, and although there is substantial cycling of sulfur between organic and inorganic
fractions, several lines of evidence suggest that there is little net retention of sulfur in the organic
fraction of most forest soils in the eastern United States. Two studies of sulfur dynamics by Fuller and
coworkers (at the Hubbard Brook Experimental Forest, New Hampshire, and the Huntington Forest,
New York), the first based on a model of kinetic data for conversion among soil sulfur pools and the
second assessing distribution of sulfur isotopes in soil sulfur pools, concluded that overall rates of
immobilization and mineralization of organic sulfur in soils were about equal and that net retention in
organic pools was near zero (Fuller et al., 1986a,b). Sulfur input-output budgets developed for the
northeastern United States as part of the DDRP (Rochelle and Church, 1987; Church et al., 1989)
show that budgets are, on average, very close to steady state, indicating that there is not significant
net retention in organics or other watershed sulfur pools on a broad regional basis. Finally, for soils in
each of the three DDRP regions, median carbon-to-sulfur ratios in mineral horizon soils were less than
160:1, and were less than 100:1 in B and C horizons. These ratios are lower than the 200:1 ratio
below which soils would be expected to have net mineralization of sulfur, and far below the 400:1 ratio
above which soil organic matter is expected to be accumulating sulfur (Barrow, 1961; T. Strickland,
pers. comm., 1989). Based on these data, it seems unlikely that soils in DDRP watersheds are
accumulating sulfur in organic matter; they might even be a net sulfur source.
10.7.4.2 Base Cation Supply
10.7.4.2.1 Cation exchange --
Baker et al., (1990) determined that mineral weathering and cation exchange were major factors
influencing surface water acid-base chemistry. The MAGIC model is based on a conceptual model
having sulfate adsorption as the principal control on acid anion mobility and neutralization of acidic
cations by cation exchange and/or mineral weathering (Galloway et al., 1983; Cosby et al., 1985a,b,c).
In M-APP simulations, the projected increase in median stream sulfate over 50 years was +57 ^eq/L,
with an associated increase in base cation (Ca + Mg) concentrations of +21 fieq/L. In the MAGIC
model simulations, mineral weathering rates are fixed during calibration, such that long-term changes
in base cations and ANC are controlled by exchange equilibrium. Thus, although ANC formulations in
MAGIC do consider aluminum dynamics, ANC can also be estimated as Z base cations - 2 acid
anions. For the constant deposition scenario, the decrease in the median ANC for M-APP projections
was about 30 fteq/L, which is consistent with the greater increase in sulfate than in base cation
concentrations.
Watershed exchange pools are tightly coupled with deposition inputs. This tight coupling of
declining base cation concentrations with declining surface water sulfate concentrations was recently
reported for the Hubbard Brook Experimental Forest (Driscoll et al., 1989b). Two mechanisms can
contribute to this coupling: (1) decreasing atmospheric deposition of base cations and sulfate and
(2) decreasing release of base cations due to reduced sulfate efflux and depletion of the exchangeable
base pool in soils (Driscoll et al., 1989b; Holdren and Church, 1989).
310
-------
A primary assumption of the DDRP is that decreased sulfur deposition rates occur in conjunction
with decreased emission rates only for sulfate and hydrogen concentrations (i.e., that concentrations
of other acid anions and base cations are exchanged). For our model simulations, electroneutrality of
deposition was maintained by decreasing hydrogen ion concentrations on a 1:1 equivalent basis with
sulfate. There currently is no good evidence that decreases in anthropogenic emissions of sulfur are
or should be necessarily accompanied by changes in base cation deposition rates (A. Olsen, pers.
comm.; R. Dennis, pers. comm.) and it remains uncertain whether base cations would decrease with
emission controls. If the major source of base cations is local entrainment and resuspension,
emission controls will have little influence on base cation concentrations in deposition. If local sources
of dust-borne base cations are reduced (i.e., by paving roads and employing dust control at construc-
tion sites) long-range transport of base cations from emissions would become an increasingly impor-
tant component of the cation/anion budget of deposition, although total cation deposition would
decrease. By ignoring changes in base cations, therefore, an additional source of uncertainty has
been included in model projections.
Process-oriented interpretation of experimental and monitoring data for watersheds in the M-APP
Region are consistent with the processes embodied in MAGIC simulations (i.e., that mineral weather-
ing, cation exchange, and sulfate sorption are the predominant controls on surface water chemistry in
the M-APP Region). Questions remain, however, as to whether there are other important watershed
processes affecting surface water acidification that should be incorporated into watershed models. It
also remains quantitatively unclear how much the model formulations, operational assumptions,
parameter selection, and data aggregation affect the results of long-term projections.
10.7.4.2.2 Mineral weathering -
Mineral weathering is critical for all long-term projections, but it is a process about which little
information at the watershed scale can be obtained. Mineral weathering parameters are calibrated as
part of the watershed-by-watershed calibration procedure used by MAGIC, but weathering rates are
not completely unconstrained. Rather, there are limits on the range over which these parameters can
vary while maintaining reasonable ranges for other, related parameters (e.g., historical base saturation)
and still matching observed surface water chemistry constituent concentrations (e.g., silica, calcium,
sodium). Acidic deposition can result in changes in soil pH (Falkengren-Grerup et al., 1987; Tamm
and Hallbacken, 1988) and changes in soil pH may influence mineral weathering rates and sulfate
absorption capacities. In the M-APP Region, however, median projected changes in 50 years for soil
solution pH were less than 0.2 units (Section 9.3). Although mineral weathering rates cannot be
unequivocally estimated, the model formulations and mass balance approaches used for the DDRP
simulations are analogous to the mass balance approaches that have been used to estimate weather-
ing in watershed studies (Velbel, 1986a; Paces, 1973).
10.7.4.3 Nitrogen Dynamics
MAGIC focuses on the effects of sulfur deposition on surface water acidification. MAGIC
considers total deposition acidity, including nitrate, but the dynamic nitrate formulations are
rudimentary. Because most eastern forested watersheds are nitrogen limited (Likens et al., 1977;
311
-------
Swank and Crossley, 1988), ft is likely that nitrogen inputs in most systems are effectively retained in
vegetation and soil, and nitrate is not an important mobile anion. Deposition inputs of nitrate are
about twice the inputs of ammonium for the eastern United States (Kulp, 1987). Although nitrification
has an acidifying effect, nitrate assimilation has an alkalizing effect (Lee and Schnoor, 1988). Nitrate
concentrations typically are low in M-APP streams and do not play a large role in surface water acid-
base status at present (Baker et al., 1990). It is possible that nitrogen might become an important
contributor to chronic surface water acidification in the future for some forested systems. Evaluation of
this possible effect was outside of the scope of this study, however. It is well known that nitrate can
be an important component of episodic acidification (Wigington et al., 1990) in some situations.
10.7.4.4 Land Use Changes
Land use changes are inevitable over a 50- to 100-year period and can have a major influence
on water quality. Aggrading forests can have an acidifying effect on watershed soils due to depletion
of soil base cations and production of organic acids in litter and root exudate (Troedsson, 1980) and
also due to increased capture of dry deposition. Agricultural land uses can also significantly affect soil
chemistry through soil amendments (lime, fertilizer), alteration of soil carbon pools, etc. The objective
of the DDRP long-term simulations is not to forecast the future but rather to provide a basis for
comparing the relative effects of alternative acidic deposition scenarios through model projections.
Omitting considerations of land use change does not seriously compromise this objective. Estimations
of land use changes are important only if absolute estimates of chemical change are required. This is
not the objective of DDRP simulations.
10.7.4.5 Organic Acidification
One of the operational assumptions of MAGIC was that the status of organic acids would remain
constant. Krug and Frink (1983) hypothesized that reversing surface water acidification by reducing
the input of strong mineral acids could result in increased dissociation of humic acids and mobility of
organic acids. This, in turn, could return the acid-base status of systems that are currently acidic (due
to mineral acidity) to a previous acidic state (due to organic acidity). Organic acid concentrations in
streams throughout the Mid-Atlantic highlands currently are low (Kaufmann et al., 1988). Because of
the nature of the watersheds and soils of the region, we believe that a reduction in acidic deposition
inputs would not lead to significantly increased leaching of organic acids to the point of causing acidic
conditions.
10.7.5 Other Considerations
10.7.5.1 Data Aggregation
In generating data at watershed scale for use in DDRP models, physical and chemical attributes
of soils were averaged (weighted) to master horizons, master horizons were aggregated to sampling
classes, and sampling class attributes were aggregated to the watershed values, which were then
used for calibration and simulations. This averaging or aggregation process preserves the central
tendency in watershed attributes and subsequent projected effects, but reduces the variability (i.e.,
312
-------
extreme values in the distribution of soil attributes at the horizon-to-watershed scale). Data
aggregation thus results in underestimates of change in the tails or extremes of the distributions.
Although these extremes represent a small proportion of the target population, the changes in these
watersheds might be underestimated, so the changes in ANC or pH for a fraction of regional surface
waters is likely to be larger than projected.
10.7.5.2 Effects on Upstream Versus Downstream Nodes
The DDRP target population represents the lower stream nodes sampled during the NSS. Very
important differences in water chemistry exist between the upstream and downstream nodes of these
reaches (Kaufmann et al., 1988). For example, the median ANC for the target population of upstream
nodes of the reaches modelled in DDRP is over 20 ^eq/L lower than that for the downstream nodes
(Table 10-14). Also, about 40 percent more upstream nodes than downstream nodes have ANC
values < 50 fieq/L. Most strikingly, over five times as many upstream nodes as downstream nodes
are acidic.
When Kaufmann et al. (1988) presented target population estimates of the lengths of stream
reaches that were acidic or had ANC < 50 ^eq/L, they used linear interpolation along the reaches.
The DDRP estimates of stream reach length projected by MAGIC to be acidic or to have ANC < 50
jueq/L are based solely on chemistry data and simulations for the lower nodes. If the DDRP estimated
stream reach length in the manner used by the NSS, the estimates of stream reach length with ANC
< 0 or 50 fteq/L would surely be much greater. For example, for the stream reaches modelled by
MAGIC, the estimate of current reach length with ANC < 50 ^eq/L based on measured NSS values for
the lower nodes is 2,505 km (Table 1O-11). Using the NSS interpolating procedure, the corresponding
estimate is 4,222 km (A. Herlihy, pers. comm.). Similarly, the estimate of acidic stream reach length is
472 km (Table 10-11), based on measured ANC at lower nodes, whereas the corresponding estimate
using the NSS interpolation is 1,557 km (A. Herlihy, pers. comm.), over three times as great.
From conducting transects of stream reach chemistry, Kaufmann et al. (1988) found that for
upper stream reaches with low ANC (like the DDRP subsample from NSS), there was a fairly smooth
and continuous decline in pH upstream from the lower node. This observation supports a hypothesis
that as watershed soils grow more shallow in the upper reaches of catchments, cumulative buffering of
soils on receiving water decreases and so, presumably, does the ability of the catchment to buffer
against the effects of acidic deposition. Thus, estimation of stream reach length < 0 or 50 /*eq/L
based solely on lower node chemistry is surely an underestimate of the length affected, as is the
estimation of future conditions based upon modelling of conditions only at the lower node. The
estimates of stream reach length projected to have ANC values < 50 /*eq/L or projected to become
acidic at some future time, as presented in this report, are probably large underestimates of
acidification or recovery as compared to projections that could be generated using interpolation
procedures analogous to those used by the National Stream Survey.
313
-------
Table 10-14. Comparison of ANC Values at Upstream and Downstream Reach Nodes for the
M-APP DDRP Modelling Target Population
Node
Median ANC («eq/L)
# sites with current ANC <
# sites with current ANC < 0 fieq/L
Downstream
87
1504
145
Upstream
64
2370
772
314
-------
10.8 REGIONAL COMPARISON: NE, M-APP, SBRP
Church et al. (1989) found that in the NE, lakes would respond relatively rapidly to changes in
sulfate deposition because the NE was, on the average, near sulfate steady state and because sulfate
retention capacity of NE soils is small. In contrast, for either current or increased sulfur deposition, it
might be 150-200 years before the SBRP watersheds are, on the average, at or near sulfate steady
state (Church et al., 1989). Most M-APP watersheds, on the average, were projected to reach sulfur
steady state in about 50 years. The M-APP Region has watersheds with characteristics of both NE
and SBRP watersheds and may represent a transition region where some systems can be regarded as
responding relatively rapidly to sulfate deposition, whereas other systems may have longer response
times. In contrast to the SBRP, which has no currently acidic streams, about 4 percent of the M-APP
stream reaches are currently acidic at their lower nodes, with acidic deposition the most likely
contributor to stream acidity (Kaufmann et al., 1988).
The MAGIC model was used to project changes in surface water chemistry in all three DDRP
regions, so relative comparisons among the three regions can be made using projections based on
simulations from a common model. In the NE, MAGIC projections indicated that less than an
additional 1 percent of the lakes might become acidic (ANC < 0 ,weq/L) in 50 years under current
deposition. This estimate represents the difference between the number of nonacidic lakes projected
to become acidic and currently acidic lakes that might become nonacidic under current deposition
levels. In the SBRP, MAGIC projections indicated 1 percent of the currently nonacidic streams might
become acidic within 50 years under current deposition and the ANC of an additional 9 percent of
streams was projected to decrease to < 50 /weq/L By contrast, in the M-APP, MAGIC projections
indicate much larger changes, with an additional 8 percent of the stream population projected to
become acidic under current deposition after 50 years, and an additional 29 percent of the stream
population with ANC decreasing to < 50 ^
With a 30 percent decrease in deposition, there might be about a 2 percent increase in the
target population of nonacidic NE lakes (i.e., currently acidic lakes might become nonacidic) and
about a 2 percent increase in the target population of NE lakes with ANC > 50 ^eq/L in 50 years. In
the M-APP, a 50 percent deposition decrease might result in no (0) acidic streams after 20-25 years
and a 3 percent increase in the target population of M-APP streams with ANC > 50 ^eq/L after 50
years. For the currently acidic streams, this represents an increase in the median ANC of about 15
^eq/L after 50 years. With an initial 50 percent decrease in deposition followed by a relative increase
to a final 20 percent deposition decrease after 50 years, there might be an increase in ANC for the first
20 years, with no acidic streams after 20 years. Following this initial ANC increase, however, there
might be increased surface water acidification with about 4 percent of the streams estimated to be
acidic after 50 years. This percent of acidic streams is comparable to the initial calibrated percent of
streams in the M-APP.
Comparing these three regions and projections with the MAGIC model for different deposition
scenarios indicates there might not be significant changes in surface water chemistry under current
deposition in the NE but that at least a 30 percent decrease in deposition might be necessary to see
an increase in surface water ANC and minimize continued surface water acidification. In the SBRP,
315
-------
some streams might become chronically acidic within 50 years under current deposition levels.
Scenarios of decreased deposition were not simulated for the SBRP. In the M-APP, a significant
percent of stream reaches might become acidic within 50 years under current deposition. A 50
percent deposition decrease might be necessary in this region to see an increase in ANC and to
minimize continued surface water acidification.
We are well aware that long-term modelling is a difficult and imperfect activity. Particularly
vexing is the problem of confirmation, both of models themselves and of their projections. Twelve-year
calibration and confirmation studies on Coweeta Watersheds 34 and 36 indicate that the RMSEs
among the models and the observed standard errors of the data are similar. Subtle long-term trends
were not effectively simulated, however, because of calibration procedures. Regardless, long-term
projections can be confirmed only with long-term periods of record (Simons and Lam, 1980), which do
not exist. Although our model projections cannot be confirmed, they are consistent with theory and
limited field experiments and observations.
10.9 CONCLUSIONS FROM LEVEL III ANALYSES
Conclusions from the Level III analyses follow:
• The M-APP Region is a transition region between NE watersheds with short response times
and SBRP watersheds with long response times.'
• Current deposition levels might result in an increase in acidic streams in the M-APP of about
threefold over 50 years, and about eightfold over 100 years.
• A 50 percent deposition decrease might result in no (0) acidic streams (acidic due to
atmospheric sulfur deposition) after 50 or 100 years.
• A 50/20 percent deposition decrease might result in an initial increase in ANC after 20-25
years followed by surface water acidification. Surface water conditions after 50 years might
be similar to initial calibrated surface water chemistry for the target population at year 0.
Some additional acidification is projected at 100 years.
• Current deposition levels might result in about a doubling of stream reaches with ANC
< 50 fieq/L (and thus susceptible to acidic episodes) at both 50 and 100 years.
• The rates of change in projected surface water acid-base chemistry using MAGIC are
comparable to observed rates of change in surface water chemistry ascribed to effects of
acidic deposition.
• A quantitative evaluation of the absolute accuracy of long-term model projections is not
available at present. Comparisons of relative projected effects among deposition scenarios is
perhaps more reliable than absolute projections.
316
-------
SECTION 11
SUMMARY OF RESULTS
This section presents an overview of the results of the DDRP analyses. Church et al. (1989)
performed statistical analyses of the interrelationships among deposition, edaphic factors, and surface
water chemistry for lake/watersheds in the NE and stream/watersheds in the SBRP. Those analyses
generally supported the postulated relationships incorporated into both the single factor models (for
sulfate adsorption and cation supply) and the integrated watershed models. For example, soil depth,
soil chemical characteristics, and watershed hydrology factors all appear as important explanatory
variables in the regressions performed. Additionally, wetlands in northeastern watersheds appear to
have an important role in influencing sulfur dynamics (see Sections 7 and 8 of Church et al., 1989).
The watersheds we studied in the M-APP, however, contain few wetland areas (see Sections 5 and 7);
thus, such effects are relatively unimportant in this region. As discussed in the Preface and in Section
8, we did not perform the lengthy statistical analyses for the watersheds of the M-APP Region.
11.1 RETENTION OF ATMOSPHERICALLY DEPOSITED SULFUR
11.1.1 Current Retention
We computed net annual retention of atmospherically deposited sulfur for watersheds not
having apparent significant sulfur internal sources (see Section 7). At present, watersheds in the NE
appear to be approximately at steady state, whereas median net retention is about 75 percent in the
SBRP. These observations are qualitatively consistent with theory (Galloway et al., 1983).
The M-APP Region is a transition zone between the NE and SBRP in terms of observed current
sulfur retention. Because of the similarities between soils in the M-APP Region and the SBRP, it is
possible that the M-APP Region at one time retained as much of the elevated sulfur deposition as is
now evident in the SBRP (i.e., 70 - 80 percent). It is also possible that sulfur deposition has
decreased this retention, perhaps very dramatically in the westernmost area (subregion 2Cn of the
National Stream Survey), which now has median percent sulfur retention of only three percent (Plate
11-1). High sulfur deposition and decreased retention, in turn, may have led to the low ANC and
acidic stream reaches (excluding stream reaches affected by acid mine drainage) identified there by
the National Stream Survey (Kaufmann et al., 1988).
11.1.2 Projected Retention
As indicated in Section 10.1.1 current sulfur retention in M-APP watersheds is highly variable,
ranging from systems already at steady state to sites with retention exceeding 70 percent. The
projected dynamic response of sulfur in M-APP watersheds is similarly variable; some watersheds have
already reached steady state while others are not projected to come to steady state under current
deposition for 150 years. The sulfate dynamics described in the remainder of this section are based
on projections using the integrated MAGIC model, but projections from the Level II sulfate modelling
(Section 9.2) are almost identical. For a scenario of constant deposition, model projections under-
estimate current sulfur retention; median simulated retention in year 0 was 26 percent (compared
317
-------
-------
Plate 11-1. Sulfur retention and wet sulfate deposition for National Surface Water Survey
subregions in the eastern United States.
318
-------
-------
NSWS SUBREGiONS
MEDIAN PERCENT SULFUR RETENTION
AND WET SULFATE DEPOSITION
2.25
yEDIAN PERCENT
SULFUR RETENTION
g o - 20
20 - 40
40 - 60
60 - 80
80 - 100
Average Annual
Wet Sulfaie ^ 2.75-
Deposiiion (g m"2 yr~')* 3.00^
3.25
Eastern lake Survey
-2.25
lledion
Subregion X Retention
1A
IB
1C
ID
IE
-H
8
-8
-12
X2.00
.00
Notional Sirsorn Survey
lledion
Subregion I Retention
2Cn
2Br,
3B
n
2As
3A
3
40
34
50
75
78
Deposition for 1980 - 1984
(A. Olseni Personal Communicat ion)
-------
-------
to 44 percent for measured retention). Retention is projected to decrease sharply during the simula-
tion period, to a median of 12 percent in 20 years, and to only 4 percent and 2 percent by years 50
and 100, respectively (Plate 11-2). Most of the projected decreases in retention occur during the first
50 years of simulations; the subsequent rate of decrease in retention (increase in sulfate concentra-
tion) slows substantially because most systems are near steady state. Median time to sulfur steady
state in the M-APP, based on Level II analyses, is considerably shorter (35 years) than in the SBRP
watersheds (61 years) characterized by Church et al. (1989).
The changes in sulfur retention represent large projected changes in sulfate concentration
during the simulation period. Levels of sulfur deposition in the M-APP are much higher than in the NE
or SBRP (Section 5.6). Steady-state sulfate concentrations in the M-APP, under a scenario of constant
deposition, are also much higher (median = 215/teq/L compared to 111 ^eq/L in the NE and 120
,aeq/L in the SBRP), and changes in concentration between current conditions and steady state are
large. Although current retention in the M-APP is lower than in the SBRP, the median change in
sulfate as systems come to steady state is larger, 107 ,weq/L versus 83 ^eq/L in the SBRP. The
maximum increase in the M-APP is almost 250 ^eq/L.
Under scenarios of reduced sulfur deposition, projected changes in sulfate are much smaller.
For the scenario with a 50 percent decrease in deposition (Scenario B), current sulfate concentrations
in most systems would be above the steady-state concentration. Concentrations are thus projected to
decrease as systems come to steady state with the lower level of deposition, from a simulated median
of 161 ^eq/L in year 0, to 134, 123, and 120 /
-------
-------
Plate 11-2. Changes in sulfur retention up to 100 years for watersheds of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 10.7.1).
320
-------
-------
Mid-Appalachian
Streams
PERCENT
SULFUR RETENTION
Model = MAGIC
3rd Quartile +
(1.5 x Interquartile Range)"
3rd Quartile
Mean
Median
1st Quartile
1st Quartile -
(1.5 x Interquartile Range)"
" Not to exceed extreme value.
Current Deposition
ct:
ce:
50 —
-50
'-100
- 100 —|
•—
S 50 —
Ul_J
" 0 -
O£L
Z3
•-1- -50 -
-100
50% Decrease
CSI IO O
50%/20% Decrease
50 -
CE:
CE:
0 —
i -50
'-100
C3 O
>- ct:
-------
-------
essentially an equilibrium process in soils, but the pool of exchangeable bases is limited. Because
most soils in watersheds considered by the DDRP are already strongly acidic, projected effects of
acidic deposition on soil pH are small (Section 9.3), and weathering rates would not be expected to
rise substantially. As long as weathering rates exceed the rate of acid anion flux from a watershed,
base cation resupply will exceed acid anion output, and systems will not become acidic. If, however,
acid anion output exceeds weathering rates, exchange can provide a temporary source of bases to
neutralize acidity; with time, however, the finite exchange reservoir will be depleted and runoff will
eventually become acidic.
Our analyses suggest that the importance of cation exchange in the control of cation concentra-
tions and ANC is greatest in systems with low surface water ANC (< 100 ^eq/L), and that in systems
with higher ANC, weathering resupply will maintain positive surface water ANC. In low ANC systems,
the relative contributions of weathering and exchange cannot be accurately determined.
11.2.2 Future Effects
Using two soil chemistry models, developed by Reuss and Johnson (1986) and by Bloom and
Grigal (1985), we assessed the potential importance of soil cation exchange in delaying surface water
acidification. These analyses represent a worst-case assessment that considers only exchange
processes in the upper two meters of soil (i.e., weathering is not considered). Projections were
generated assuming sulfate to be completely mobile. Model projections reflect two processes occur-
ring within the soil: (1) a mobile anion effect (i.e., to maintain charge balance, concentrations of both
basic and acidic cations increase or decrease with changes in sulfate) and (2) depletion of the soil
base cation exchange pool (i.e., as the proportion of base cations on the solid phase is depleted, the
proportion of bases in solution will also decrease and the proportion of acidic cations in solution will
increase). Results of simulations from the two cation models are generally consistent. For a
deposition Scenario A, of constant deposition, ANC is projected to decrease only slightly during the
first 50 years of simulations, then to drop sharply between years 50 and 100 as the exchange pool
becomes depleted (median projected decreases 0.3, 3, and ys^weq/L at 20, 50, and 100 years). For
the same period, soil base saturation is projected to decline steadily, from 14 to 9 percent; soil pH
drops slightly during the first 50 years of simulations, but the decline then accelerates as the
exchange buffer system is depleted. Soil solution aluminum concentrations increase in a corre-
sponding nonlinear pattern, from 9 ^mol/L at the start of simulations to 22 jumol/L at year 100. For
the scenarios of reduced sulfur deposition, changes are much smaller. ANC is projected to increase
(median increase of 6 ^aeq/L after 100 years) as a consequence of deposition Scenario B, due to a
mobile anion effect. Soil pH, base saturation, and soil solution aluminum remain almost constant
under the same scenario. For Scenario C, ANC is projected to initially increase, then to decrease by a
median of 8 ^eq/L after 100 years. For the same scenario, soil pH and base saturation decrease over
the 100-year simulation period, but by much smaller amounts than if current rates of deposition are
maintained.
321
-------
11.3 INTEGRATED EFFECTS OF SURFACE WATER ANC
As noted in the Preface of this report, as discussed in Section 1.3.4, and as fully described in
Section 10, we applied the MAGIC watershed model to study the stream watersheds in the M-APP
Region. At present, it is not possible to quantify the long-term (i.e., 50-100 years) accuracy of future
projections of any watershed acidification model, including the MAGIC model. Calibration and short-
term (10 years) verification tests, as well as comparisons to other field and laboratory studies and
paleoecological reconstructions of past lake chemistries, however, are reasonably encouraging (see
Sections 10 and 1.5).
The MAGIC model was successfully calibrated for 29 of the 36 DDRP stream reaches (at the
lower nodes; see Section 10) in the M-APP. The 36 DDRP watersheds represent 5,496 stream
reaches with a combined reach length of 15,239 km. The 29 watersheds simulated by MAGIC
represent 4,298 stream reaches totalling 11,246 km of stream reach length. Modelling estimates of
effects on stream chemistry were performed for time periods up to 100 years. As discussed in Section
10, the modelling of the stream reaches, based on calibrations to observed chemical concentrations at
the lower nodes of each stream reach, resulted in projections of conditions for the most buffered part
(i.e., the lower end) of the stream reaches under study (see also Kaufmann et al., 1988). Therefore,
projections of changes in chemical concentrations for stream reach lengths are conservative (i.e.,
underestimates).
The fact that calibrated sulfate concentrations exceed observed concentrations also has an
effect (see Table 10-6). For projections of increased sulfate concentration, the change from the
calibrated value to the steady-state value is an underestimate, leading to an underestimate of acidi-
fication (i.e., loss of ANC). For projections of decreased sulfate concentrations, the effect is an
overestimate of ANC recovery.
As discussed in Section 10, the watershed modelling analyses make use of watershed soil
representations as aggregated from the DDRP Soil Survey. Because the focus of the DDRP is on
regional characteristics and responses, soils data were gathered and aggregated to capture the most
important central tendencies of the systems under study. As a result, extremes of individual watershed
responses are underestimated in the analyses presented in this report. Those systems that MAGIC
projects to respond to the greatest extent or most quickly to current or altered levels of sulfur
deposition might, in fact, respond even more extensively or more quickly than indicated here. This
should be kept in mind when evaluating the results of the MAGIC simulations.
The projected changes in the acid-base status of M-APP streams can be presented in a variety
of ways. Plate 11-3 illustrates the projected changes in the distributions of ANC for each of the three
deposition scenarios for up to 100 years. Plate 11-4 shows the associated projected changes in pH.
More detail on the projected changes in ANC and pH at 50 years is shown in Plates 11-5 and 11-6,
respectively. Plates 11-3 and 11-4 indicate patterns of (1) a monotonic decrease in both ANC and pH
for the constant deposition scenario, and (2) an increase in ANC and pH in the short term (20 years),
with longer term decreases for both decreased deposition scenarios. For the 50 percent decreased
deposition scenario, there are slight decreases in ANC and pH over the long term, presumably due to
322
-------
Plate 11-3. Changes In ANC up to 100 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 10.7.1).
323
-------
-------
Mid-Appalachian
Streams
ANC vs. TIME
Model = MAGIC
300
200
100
0
•100
Current Deposition
Cvl UO
CtZ C£.
r.
3rd Quoriiie +
(1.5 x Interquartile Range)"
3rd Quortile
Mean
Uedian
1st Quorlile
1st Quartile -
(1.5 x Interquartile Range)"
" Not to exceed extreme value.
300 —i
< 200 -
CT
.£100-
= 0 H
-100 -
50% Decrease
o
az.
Cvl
oz
300
^L 200
cr
a>
=> 100
50%/20% Decrease
<-3
-too
CvJ LO
CtZ &Z
o
<=s
-------
-------
Plate 11-4. Changes in pH up to 100 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 10.7.1).
324
-------
-------
Mid-Appalachian
Streams
pH vs. TIME
Model = MAGIC
8.0
7.0
6.0
5.0
4.0
Current Deposition
CV1 LO O
Ct± •«—
>— Ct: CrZ
>- >— ct:
3rd Quortile +
(1.5 x Interquartile Range)"
3rd Quartile
Mean
Median
1st Quartile
1st Quartile -
(1.5 x Interquartile Range)"
" Not to exceed extreme value.
8.0-
7.0 -
5,0 —
5.0 -
4.0-
50% Decrease
c-j to
8.0
7.0
6.0
5. 0
4.0
50%/20% Decrease
O C3
O£
-------
-------
Plate 11-5. Changes in median ANC at 50 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 10.7.1).
325
-------
-------
CHANGE IN MEDIAN ANC
Year 0 to Year 50
Model = MAGIC
l-APP Study Area
C3T>
c=
C3
30
20
10H
Constant
Deposition
20-
30-
50%
Decrease
30n
20
10-I
rr~
1 0-
20-
30-
is
:» j
50%/20%
Decrease
30n
20-
10-
0
10
20
30
10
UO
V
i i
tO CD
V
to
to cz>
OJ CZ>
ANC Group (ueq/L)
-------
-------
Plate 11-6. Changes in median pH at 50 years for stream reaches of the M-APP Region as
projected by the MAGIC model for its specified DDRP target population (see Section 10.7.1).
326
-------
-------
CHANGE IN MEDIAN pH
Year 0 to Year 50
Model = MAGIC
I-APP Study Area
er-
as
CTl
c=
1 ,0
0 ,5H
0
Constant
Deposition
50%
Decrease
o ,5
-0,40
-0.11
-0,29
IO CD CD
OM CD CD
\S
i i
LO CD
1 .0
0,5
0
-0.5H
0.67
0.05
°-09
Cvj
\s
CD CD
CD CD
I I
IO CD
Osl CD
ANC Group (ueq/L)
50%/20%
Decrease
1.0-1
0,5-
0
0.5H
0.05
-0.05 -0.02
to o
CM C3>
OJ CD
-------
-------
modelled depletion of the soil cation exchange complex. For the 50/20 percent decrease deposition
scenario, the decreases are due principally to the ramped increase of deposition back up to the 20
percent level at year 50. Plates 11-5 and 11-6 show that, although the greatest ANC changes are not
projected for the lowest ANC group (i.e., < 25 fieq/L), the largest pH changes are projected for this
group. This reflects, of course, the logarithmic relationship between these two variables.
Another way to present the results of the modelling is as changes in the percent or length of
stream reaches projected to have ANC < Q /teqfL (Figures 11-1 and 11-2) or < 50,weq/l_ (Figures 11-3
and 11-4) in the future. Because MAGIC has some bias in its calibration to year 0 (see Section 10),
these values are computed by adding the projected changes in ANC to the values observed by
Kaufmann et al. (1988) for the National Stream Survey. The changes in ANC projected by MAGIC are
unbiased with respect to year-0 ANC (J. Cosby, pers. comm.). A one-sided hypothesis test of
significance for the changes in numbers of systems (i.e., different from zero; see Section 6) indicates
changes (p < .05) at years 20, 50, and 100 for stream reaches with ANC < 0 fieq/L for the current
deposition scenario (see Section 10). Changes at 50 and 100 years for the projections of systems
with ANC < 50 /ieq/L are also evident (p = .01). For the 50/20 percent decreased deposition
scenario, changes (increases) in the numbers of stream reaches with ANC < 0 or ANC < 50 yeq/L are
indicated (.07 - p ^ .09) at both 50 and 100 years. All other projected changes are less well
supported (p > .16) (see Table 10-9).
Considered together, the results shown in Plates 11-3 through 11-6 and Figures 11-1 through
11-4 indicate a modelled projection of marked decreases in the ANC and pH of M-APP stream reaches
(as represented by chemical conditions at the lower reach nodes) under continued levels of current
deposition. These projected changes in ANC translate also to marked increases in (1) the number of
acidic stream reaches (ANC < 0 ^weq/L), and (2) the number of stream reaches with ANC < 50 jueq/L.
Stream reaches with ANC < 50 /ueq/L are probably highly susceptible to episodic acidification to
acidic conditions during periods of heavy snowmelt or rainfall (Eshleman, 1988).
Projections for the 50 percent decreased deposition scenario indicate somewhat improved (rela-
tive to toxic conditions for fish) chemical conditions at 50 years (especially pH for low ANC systems),
with perhaps a slight decrease in water quality after that time. Projections for the 50/20 percent
decreased deposition scenario indicate chemical conditions at 50 years very close to those observed
by Kaufmann et al. (1988) in the National Stream Survey. Decreases in water quality are projected at
100 years under this deposition scenario.
11.4 SUMMARY DISCUSSION
Our analyses of (1) current sulfur mobility, (2) projected sulfur mobility, (3) cation exchange
buffering, and (4) integrated watershed modelling present an internally consistent picture of the
potential long-term effects of sulfur deposition on the chemistry of our target population of stream
reaches in the M-APP Region. Although soils of our study watersheds have greater ability to retain
future atmospheric inputs of sulfate than do those of the NE, this ability is quite limited relative to soils
of the SBRP (Church et al., 1989). Soils of the M-APP Region are projected to reach sulfur input-
output steady state faster than those of the SBRP (median time to steady state is 35 years for the
327
-------
d>
30
20-
c
0>
o
o> 10
D.
Current Deposition ^
eo
Y////A
20
50
100
30
O
D)
« 20
C
o
10-
50% Decreased Deposition
CO
Y////A
0
20
50
100
30
0
O)
« 20-
c
a)
o
o
DL
50
%/20% Decreased Deposition
CO
CO *
"7 S~7 S~7 7
X/////A ° '/////,
•
0 20 50 100
Time (yr)
Figure 11-1. F3ercentage of stream reaches of the DDRP M-APP modelling target population
(see Section 10.7.1) projected by the MAGIC model to have ANC < 0 fieq/L at the
lower reach nodes. Estimates at year 0 are from the National Stream Survey
(Kaufmann et al., 1988) for the same target population. Estimates at 20, 50, and
100 years are computed by adding the change in ANC projected by the MAGIC
model to ANC observed by the National Stream Survey.
328
-------
,
c
-------
100
§,
CO
50-
-------
D)
C
-
r-
CM
20
50
co
100
5
o
(0
o
CC
10000 -
7500-
5000-
2500-
rt .
50%
m
o
in
CM
HI
Decreased Deposition
t~
CM
CM
CM
^
CM 2
S S
Wk HI
20
50
100
10000
D)
C
0>
JC
o
C3
o
CC
7500-
5000-
2500-
0
50%/20% Decreased Deposition
O)
CD
in
o
m
CM
CM
CM
CM
0 20 50 100
Time (yr)
Figure 11-4. Length (km) of stream reaches of the DDRP M-APP modelling target popula-
tion (see Section 10.7.1) projected by the MAGIC model to have ANC < 50
fieq/L at the lower reach nodes. Estimates at year 0 are from the National
Stream Survey (Kaufmann et al., 1988) for the same target population. Esti-
mates at 20, 50, and 100 years are computed by adding the change in ANC
projected by the MAGIC model to ANC observed by the National Stream
Survey.
331
-------
M-APP, 61 years for the SBRP). Calculated steady-state sulfate concentrations in the M-APP Region
(214
-------
SECTION 12
REFERENCES
Abrahamsen, G., J. Hoviand, and S. Hagver. 1980. Effects of artificial acid rain and liming on soil
organisms and the decomposition of organic matter, pp. 341-362. In: T.C. Hutchinson and M.
Havas, eds. Effects of Acid Precipitation on Terrestrial Ecosystems. Plenum Press, New York.
Adams, F., and Z. Rawajfih. 1977. Basaluminite and alunite: A possible cause of sulfate retention by
acid soils. Soil Sci. Soc. Am. J. 41:686-692.
Altshuller, A.P., and R.A. Linthurst. 1984. The Acidic Deposition Phenomenon and Its Effects: Critical
Assessment Review Papers. EPA/600/8-83/016bf. U.S. Environmental Protection Agency,
Washington, DC.
Asbury, C.E., F.A. Vertucci, M.D. Mattson, and G.E. Likens. 1989. Acidification of Adirondack lakes.
Environ. Sci. Technol. 23:362-365.
Aylmore, L.A., G.M. Karim, and J.P. Quirk. 1967. Adsorption and desorption of sulfate ions by soil
constituents. Soil Sci. 103:10-15
Baker, L.A., and P.L. Brezonik. 1988. Dynamic model of in-lake alkalinity generation. Water Resour.
Res. 24:65-74.
Baker, LA., P.L. Brezonik, and E.S. Edgerton. 1986a. Sources and sinks of ions in a softwater acidic
lake in Florida. Water Resour. Res. 22:715-722.
Baker, L.A., P.L. Brezonik, and C.D. Pollman. 1986b. Model of internal alkalinity generation: Sulfate
retention component. Water Air Soil Pollut. 31:89-94.
Baker, LA., P.R. Kaufmann, A.T. Herlihy, and J.M. Eilers. 1990. Current Status of Surface Water Acid-
Base Chemistry. NAPAP SOS/T Report 9, In: Acid Deposition: State of Science and Technology.
National Acid Precipitation Assessment Program, Volume II, Washington, DC.
Baker, L.A., P.R. Herlihy, P.R. Kaufmann, and J.M. Eilers. 1991. Acidic lakes and streams in the United
States: The role of acidic deposition. Science. 252:1151-1154.
Bard, Y. 1974. Nonlinear Parameter Estimation. Academic Press, New York. 341pp.
Barrow, N.J. 1961. Studies on mineralization of sulfur from soil organic matter. Aust. J. Agr. Res.
12:306-319.
Barrow, N.J. 1967. Studies on the adsorption of sulfate by soils. Soil Sci. 104:342-349.
333
-------
Bartling, M.H., M.L. Papp,, and R.D. Van Remortel. 1988. Direct/Delayed Response Project:
Preparation Laboratory Standard Operating Procedures. In: M.L Papp and R.D. Van Remortel. in
prep. Direct/Delayed Response Project: Preparation Laboratory Operations and Quality
Assurance Report for the Mid-Appalachian Soil Survey.
Bayley, S.E., R.S. Behr, and C.A. Kelly. 1986. Retention and release of S from a freshwater wetland.
Water Air Soil Pollut. 31:101-114.
Bettany, J.R., J.W.B. Stewart, and E.H. Halstead. 1973. Sulfur fractions and carbon, nitrogen and sulfur
relationships in grasslands, forest and associated transitional soils. Soil Sci. Soc. Am. J.
37:915-918.
Beven, K.J., and J.M. Kirkby. 1979. A physically based, variable constant contributing area model of
basin hydrology. Hydrol. Sci. Bull. 24:43-69.
Bloom, P.R., and DP. Grigal. 1985. Modeling soil response to acidic deposition in nonsulfate
adsorbing soils. J. Environ. Qual. 14:489-495.
Bornemisza, E., and R. Llanos. 1967. Sulfate movement, adsorption, and desorption in three Costa
Rican soils. Soil Sci. Soc. Am. Proc. 31:356-360.
Brezonik, P.L., L.A. Baker, and T.E. Perry. 1987. Mechanisms of alkalinity generation in acid-sensitive
softwater lakes, pp. 229-262. In: R. Hites and S.J. Eisenreich, eds. Sources and Fates of Aquatic
Pollutants. Advances in Chemistry, Series 216. American Chemical Society, Washington, DC.
Byers, G.E., R.D. Van Remortel, M.J. Miah, J.E. Teberg, M.L. Papp, B.A. Schumacher, B.L. Conkling,
D.L. Cassell, and P.W. Shaffer. 1990. Direct/Delayed Response Project: Quality Assurance
Report for Physical and Chemical Analyses of Soils from the Mid-Appalachian Region of the
United States. EPA/600/4-90/001. U.S. Environmental Protection Agency, Washington, DC.
337 pp.
Campbell, W.G., and M.R. Church. 1989. EPA uses GIS to study lake and stream acidification. Federal
Digital Cartography Newsletter 9:1-2.
Campbell, W.G., M.R. Church, G.D. Bishop, D.C. Mortenson, and S.M. Pierson. 1989. The role for a
Geographic Information System in a large environmental project. Internal. J. GIS. 3:349-362.
Chan, W.H., D.B. Orr, and R.J. Vet. 1982a. The Acid Precipitation in Ontario Study (APIOS). An
Overview: The Cumulative Wet/Dry Deposition Network. Report #ARB-15-82-ARSP, Ontario
Ministry of the Environment.
Chan, W.H., D.B. Orr, and R.J. Vet. 1982b. The Acid Precipitation in Ontario Study (APIOS). An
Overview: The Event Wet/Dry Deposition Network. Report #ARB-11-82-ARSP, Ontario Ministry of
the Environment.
334
-------
Chang, J.S., R.A. Brost, I.S.A. Isaksen, S. Modronich, P. Middleton, W.R. Stockwell, and C.J. Walcek.
1987. A three-dimensional Eulerian acid deposition model: Physical concepts and formulation.
J. Geophys. Res. 92:14681-14700.
Chao, T.T., M.E. Harward, and S.C. Fang. 1962a. Movement of ^S tagged sulfate through soil
columns. Soil Sci. Soc. Am. Proc. 26:27-32.
Chao, T.T., M.E. Harward, and S.C. Fang. 1962b. Adsorption and desorption phenomena of sulfate
ions in soils. Soil Sci. Soc. Am. Proc. 26:234-237.
Chao, T.T., M.E. Harward, and S.C. Fang. 1964a. Anionic effects on sulfate adsorption by soils. Soil
Sci. Soc. Am. Proc. 28:581-583.
Chao, T.T., M.E. Harward, and S.C. Fang. 1964b. Iron or aluminum coatings in relation to sulfate
adsorption characteristics of soils. Soil Sci. Soc. Am. Proc. 28:632-635.
Charles, D.F., and J.P. Smol. 1990. The PIRLA II Project: Regional assessment of lake acidification
trends. Verh. Internat. Verein. Limnol. 24:474-480.
Chen, C.W., S. Gherini, R.J.M. Hudson, and J.D. Dean. 1983. The Integrated Lake/Watershed
Acidification Study, Volume 1: Model and Procedures. Rep. No. EA-3221. Electric Power
Research Institute, Palo Alto, CA.
Chen, C.W., S.A. Gherini, R.J.M. Hudson, and J.D. Dean. 1984. The Integrated Lake-Watershed
Acidification Study, Volume 2: Hydrologic Analysis. C.W. Chen, ed. Rep. No. EA-3221. Electric
Power Research Institute, Palo Alto, CA.
Chou, L, and R. Wollast. 1985. Steady-state kinetics and dissolution mechanisms of albite. Am. J. Sci.
285:963-993.
Chrisman, N.R. 1987. Efficient digitizing through the combination of appropriate hardware and software
for error detection and editing. Intl. J. Geog. Info. Systems. 1:265-277.
Church, M.R. 1989. Predicting the future long-term effects of acidic deposition on surface water
chemistry: The Direct/Delayed Response Project. Eos, Transactions, American Geophysical Union
70:801-802,812-813.
Church, M.R., and R.S. Turner. 1986. Factors Affecting the Long-term Response of Surface Waters to
Acidic Deposition: State of the Science. EPA/600/3-86/025. NTIS PB86 178 188/AS. U.S.
Environmental Protection Agency, Corvallis, OR. 274 pp.
335
-------
Church, M.R., K.W. Thornton, P.W. Shaffer, D.L Stevens, B.P. Rochelle, G.R. Holdren, M.G. Johnson,
JJ. Lee, R.S. Turner, D.L Cassell, DA Lammers, W.G. Campbell, C.I. Liff, C.C. Brandt, LH. Liegel,
G.D. Bishop, D.C. Mortenson, S.M. Pierson and D.D. Schmoyer. 1989. Direct/Delayed Response
Project: Future Effects of Long-term Sulfur Deposition on Surface Water Chemistry in the
Northeast and Southern Blue Ridge Province. EPA/600/3-89/061 a-d, U.S. Environmental
Protection Agency, Washington, DC, 887 pp.
Clark, T.L, R.L. Dennis, and S.K. Seilkop. 1989. Re-examination of Interim Estimates of Annual Sulfur Dry
Deposition Across the Eastern United States. EPA/600/4-89/026. U.S. Environmental Protection
Agency, Atmospheric Research and Exposure Assessment Laboratory, Research Triangle Park,
NC, 34 pp.
Cochran, W.G. 1977. Sampling Techniques, pp. 259-261. John Wiley & Sons, Inc., New York, NY.
Comfort, S.D., R.P. Dick, and J. Baham. 1991. Air-drying and pretreatment effects on soil sulfate
sorption. Soil Sci. Soc. Am. J. 55:968-973.
Cook, R.B., and D.W. Schindler. 1983. The biogeochemistry of sulfur in an experimentally acidified lake.
Ecol. Bulletin (Stockholm). 35:115-127.
Cosby, J., University of Virginia, Charlottesville, Virginia, personal communication.
Cosby, B.J., G.M. Hornberger, J.N. Galloway, and R.F. Wright. 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.
Cosby, B.J., G.M. Hornberger, J.N. Galloway, and R.F. Wright. 1985b. Time scales of catchment
acidification: A quantitative model for estimating freshwater acidification. Environ. Sci. Technol.
19:1144-1149.
Cosby, B.J., G.M. Wright, G.M. Hornberger, and J.N. Galloway. 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.
Cosby, B.J., G.M. Hornberger, E.B. Rastetter, J.N. Galloway, and R.F. Wright. 1986a. Estimating
catchment water quality response to acid deposition using mathematical models of soil ion
exchange processes. Geoderma 38:77-95.
Cosby, B.J., G.M. Hornberger, R.F. Wright, and J.N. Galloway. 1986b. Modeling the effects of acid
deposition: Control of long-term sulfate dynamics by soil sulfate adsorption. Water Resour. Res.
22:1283-1292.
Cosby, B.J., P.G. Whitehead, and R. Neale. 1986c. A preliminary model of long-term change in stream
acidity in southwest Scotland. J. Hydro!. 84:381-401.
336
-------
Cosby, B.J., G.M. Hornberger, and R.F. Wright. 1987. A regional model of surface water acidification in
southern Norway: calibration and validation using survey data. In: J. Kamari, ed. Proceedings of
the IISAS-IMGW Task Force Meeting on Environmental Impact Models to Assess Regional
Acidification. Reidel Publishing.
Cosby, B.J., G.M. Hornberger, P.P. Ryan, and D.M. Wolock. 1989. MAGIC/DDRP Final Report. Volume
1. Models, Calibration, Results Uncertainty Analysis, QA/QC. U.S. EPA Environmental Research
Laboratory, Corvallis, OR. 286 pp.
Cosby, B.J., A. Jenkins, R.C. Ferrier, J.D. Miller, and TAB. Walker. 1990. Modelling stream acidification
in afforested catchments: Long-term reconstructions at two sites in Central Scotland. J. Hydrol.
120:143-162.
Cosby, B.J. and P.P. Ryan. 1991. Effects of inorganic and organic constituents on the long-term
variation of the acid-base status of small lakes. Presented °at the Symposium on Environmental
Chemistry of Lakes and Reservoirs, American Chemical Society National Meeting, April 16-17,
Atlanta, GA.
Couto, W., D.J. Lathwell, and D.R. Bouldin. 1979. Sulfate sorption by oxisols and an alfisol of the tropics.
Soil Sci. 127:108-116.
Cozzarelli, I.M. 1986. Processes of aluminum mobilization and immobilization in a natural soil system,
White Oak Run, Virginia. M.S. Thesis, Department of Environmental Sciences, University of
Virginia, Charlottesville, VA.
Cronan, C.S., and C.L Schofield. 1979. Aluminum leaching response to acid precipitation: Effects on
high-elevation watersheds in the Northeast. Science 204:304-306.
Gumming, B.F., J.P. Smol, J.C. Kingston, D.F. Charles, H.J.B. Birks, K.E. Cambun, S.S. Dixit,
A.J. Uutala, and A.R. Selle. How much acidification has occurred in Adirondack region lakes
(New York, USA) since pre-industrial times? Can. J. Fish. Aquat. Sci. (In press).
Dahlgren, R.A., D.C. McAvoy, and C.T. Driscoll. 1990. Acidification and Recovery of a Spodosol Bs
Horizon from Acidic Deposition. Environmental Science and Technology. 24:531-537.
Dasch, J.M. 1987. Measurement of dry deposition to surfaces in deciduous and pine canopies. Environ.
Pollut. 44:261-277.
David, M.B., and M.J. Mitchell. 1985. Sulfur constituents and cycling in waters, seston, and sediments of
an oligotrophic lake. Limnol. Oceanogr. 30:1196-1207.
David, M.B., M.J. Mitchell, and J.P. Nakas. 1982. Organic and inorganic sulfur constituents of a forest
soil and their relationship to microbial activity. Soil Sci. Soc. Am. J. 46:847-852.
337
-------
David, M.B., M J. Mitchell, and S.C. Schindler. 1984. Dynamics of organic and inorganic sulfur
constituents in hardwood forest soils, pp. 221-246. In: E.L Stone, ed. Forest Soils and Treatment
Impacts: Proceedings of the Sixth North American Forest Soils Conference, June 1983. Dept.
Forestry, Fisheries, and Wildlife, University of Tennessee, Knoxville.
David, M.B., M.J. Mitchell, and T.J. Scott. 1988. Importance of biological processes in the sulfur budget
of northern hardwood ecosystem. Biol. Fert. Soils. 5:258-264.
Dennis, R., National Oceanic and Atmospheric Administration, Research Triangle Park, North Carolina,
personal communication.
Dillon, P.J., R.A. Reid, and R. Girard. 1986. Changes in the chemistry of lakes near Sudbury, Ontario,
following reductions of SC^ emissions. Water Air and Soil Pollut. 31:59-65.
Driscoll, C., Syracuse University, Syracuse, New York, personal communication.
Driscoll, C.T., and R.M. Newton. 1985. Chemical characteristics of Adirondack lakes. Environ. Sci.
Technol. 19:1018-1024.
Driscoll, C.T., C.P. Yatsko, and F.J. Unangst. 1987a. Longitudinal and temporal trends in the water
chemistry of the North Branch of the Moose River. Biogeochemistry 3:37-61.
Driscoll, C.T., B.J. Wyskowski, C.C. Consentini, and M.E. Smith. 1987b. Processes regulating temporal
and longitudinal variations in the chemistry of a low-order woodland stream in the Adirondack
Region of Mew York. Biogeochemistry 3:225-241.
Driscoll, C.T., N.M. Johnson, G.E. Likens, and M.C. Feller. 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.
Driscoli, C.T., R.D. Fuller, and W.P. Schecher. 1989a. The role of organic acids in the acidification of
surface waters in the eastern U.S. Water Air Soil Pollut. 43:21-40.
Driscoll, C.T., G.I-.. Likens, LO. Hedin, J.S. Easton, and F.H. Bormann. 1989b. Changes in the chemistry
of surface waters: 25-year results at the Hubbard Brook Experimental Forest, NH. Environ. Sci.
Technol. 23:137-143.
Driscoll, C.T., M. Lehtinen, R. Munson, and S.A. Gherini. 1990. The contribution of strong inorganic
acids and organic acids in regulating hf and Al in Adirondack, NY (USA) Lakes. Presented at the
International Conference on Acidic Deposition. Its Nature and Impacts. Glasgow, Scotland.
September 16-21.
338
-------
Eder, B.K., and R.L. Dennis. 1990. On the use of scavenging ratios for the inference of surface-level
concentrations and subsequent dry deposition of Ca2*, Na+, and K1". Water Air Soil Pollut.
52:197-216.
Eder, B.K., D.H. Coventry, T.L. Clark, and C.E. Bellinger. 1986. RELMAP: A Regional Lagrangian Model
of Air Pollution User's Guide. EPA/600/8-86/013. NTIS PB86 171 394/AS. U.S. Environmental
Protection Agency, Research Triangle Park, NC.
Eshleman, K.N. 1988. Predicting regional episodic acidification of surface waters using empirical models.
Water Resour. Res. 24:1118-1126.
ESRI. 1987. ARC/INFO Users Guide. Environmental Systems Research Institute, Redlands, CA.
Eyre, F.H. 1980. Forest cover types of the United States and Canada. Society of American Foresters,
Washington, DC.
Falkengren-Grerup, U. 1987. Long-term changes in pH of forest soils in southern Sweden. Environ.
Pollut. 43:79-90.
Fernandez, I.J., and P.A. Kosian. 1987. Soil air carbon dioxide concentrations in a New England
spruce-fir forest. Soil Sci. Soc. Am. J. 51:261-263.
Fisher, R.A. 1935. The logic of inductive inference (with discussion). J. Royal Stat. Soc. 98:39-54.
Fitzgerald, J.W., and D.W. Johnson. 1982. Transformations of sulphate in forested and agricultural lands,
pp. 411-426. In: A.I. Moore, ed. Sulphur, Volume 1. British Sulphur Corp., London.
Fitzgerald, J.W., and M.E. Watwood. 1987. Sulfur retention in hardwood forest litter and soil: Mechanisms
and environmental parameters, pp. 119-126. In: NAPAP Aquatic Effects Task Group VI Peer Review
Project Summaries, Volume I, May 17-23. New Orleans, LA.
Fitzgerald, J.W., T.C. Strickland, and W.T. Swank. 1982. Metabolic fate of inorganic sulphate in soil
samples from undisturbed and managed forest ecosystems. Soil Biol. Biochem. 14:529-536.
Fletcher, R., and M.J.D. Powell. 1963. A rapid convergent descent method for minimization. Computer J.
6:163-168.
Fuller, R.D., M.B. David, and C.T. Driscoll. 1985. Sulfate adsorption relationships in forested Spodosols of
the northeastern USA. Soil Sci. Soc. Am. J. 49:1034-1040.
Fuller, R.D., C.T. Driscoll, S.C. Schindler, and M.J. Mitchell. 1986a. A simulation model of sulfur
transformations in forested Spodosols. Biogeochemistry 2:313-328.
339
-------
Fuller, R.D., MJ. Mitchell, H.R. Krouse, B.J. Wyskowski, and C.T. Driscoll. 1986b. Stable sulfur isotope
ratios as a tool for interpreting ecosystem sulfur dynamics. Water Air Soil Pollut. 28:163-171.
Gaines, Jr., G.J., and H.C. Thomas. 1953. Adsorption studies on clay minerals. II. A formulation of the
thermodynamics of exchange adsorption. J. Chem. Phys. 21:714-718.
Galloway, J.N., S.A. Norton, and M.R. Church. 1983. Freshwater acidification from atmospheric
deposition of sulfuric acid: A conceptual model. Environ. Sci. Technol. 17:541-545.
Galloway, J.N., G.R. Hendrey, C.L. Schofield, N.E. Peters, and A.H. Johannes. 1987. Processes and
causes of lake acidification during spring snowmelt in the west-central Adirondack Mountains, New
York. Can. J. Fish. Aquat. Sci. 44:1595-1602.
Gherini, S.A., L Mok, R.J. Hudson, G.F. Davis, C.W. Chen, and R.A. Goldstein. 1985. The ILWAS model:
Formulation and application. Water Air Soil Pollut. 26:425-459.
Gobran, G.R., and S.I. Nilsson. 1988. Effects of forest floor leachate on sulfate retention in a Spodosol
soil. J. Envir. Qual. 17:235-239.
Goldstein, R.A., and S.A. Gherini, eds. 1984. The Integrated Lake-Watershed Acidification Study. Volume
4: Summary of Major Results. EA-3221, Research Project 1109-5. Electric Power Research
Institute, Palo Alto, CA.
Graczyk, D., U.S. Geological Survey, Madison, Wisconsin, personal communication.
Grennfelt, P. 1987. Deposition processes for acidifying compounds. Environ. Technol. Letters 8:515-527.
Grieve, I.C. 1989. A Laboratory test of the soil chemical submodels of two models of catchment
acidification. Hydrological Processes. 3:339-346.
Gschwandtner, G., K.C. Gschwandtner, and K. Eldridge. 1985. Historic emissions of sulfur and nitrogen
oxides in the United States from 1900 to 1980. EPA/600/7-85/009a. U.S. Environmental Protection
Agency, Research Triangle Park, NC. 105 pp.
Haines, T.A. 1981. Acidic precipitation and its consequences for aquatic ecosystems: A review. Trans.
Am. Fish. Soc. 110:669-707.
Haque, I., and D. Walmsley. 1973. Adsorption and desorption of sulfate in some soils of the West Indies.
Geoderma 9:269-278.
Hayden, N.J. 1987. Sulfate Retention and Release in Six Eastern Soils. M.S. Thesis. Dept. Civil and
Environ. Engineering, Michigan State University, East Lansing, Ml. 165 pp.
340
-------
Hedin, L.O., G.E. Likens, K.M. Postek, and C.T. Driscoll. 1990. A field experiment to test whether
organic acids buffer acid deposition. Nature. 345:798-800*
Helgeson, H.C., W.M. Murphy, and P. Aagaard. 1984. Thermodynamic and kinetic constraints on
reaction rates among minerals and aqueous solutions: II. Rate constants, effective surface area
and the hydrolysis of feldspar. Geochim. Cosmochim. Acta 48:2405-2432.
Helvey, D., USDA Forest Service, Parsons, West Virginia, personal communication.
Henriksen, A., and D.F. Brakke. 1988. Increasing contributions of nitrogen to the acidity of surface waters
in Norway. Water Air Soil Pollut. 42:183.
Herlihy, A., Oregon State University, Corvallis, Oregon, personal communication.
Hicks, B., National Oceanic and Atmospheric Administration, Washington, D.C., personal communication.
Hicks, B.B., R.P. Hosker, Jr., and J.D. Womack. 1986. Comparisons of wet and dry deposition derived
from the first year of trial dry deposition monitoring, pp. 486-490. In: Symposium on Acid Rain: I.
Sources and Atmospheric Processes. Presented before the Division of Petroleum Chemistry, Inc.
American Chemical Society, April 13-18, New York, NY.
Hingston, F.L, R.J. Atkinson, A.M. Posner, and J.P. Quirk. 1967. Specific adsorption of anions. Nature
215:1459-1461.
Hingston, F.J., A.M. Posner, and J.P. Quirk. 1972. Anion adsorption by geothite and gibbsite: I. The role
of the proton in determining adsorption envelopes. Soil Sci. 23:177-192.
Holdren, G.R., and P.M. Speyer. 1985. pH dependent changes in the rates and stoichiometry of
dissolution of an alkali feldspar at room temperature. Am. J. Sci. 285:994-1026.
Holdren, G.R., and M.R. Church. 1989. Discussion of Driscoll, C.T., G.E. Likens, L.O. Hedlin, J.S. Eaton,
and F.H. Bormann. "Changes in the chemistry of surface waters." Environ. Sci. Technol. 23:137-
143.
Hornberger, G.M., K.J. Beven, B.J. Cosby, and D.E. Sappington. 1985. Shenandoah Watershed Study:
Calibration of a topography-based, variable contributing area hydrological model to a small
forested catchment. Water Resour. Res. 21:1841-1850.
Hornberger, G.M., B.J. Cosby, and J.N. Galloway. 1986. Modeling the effects of acid deposition:
Uncertainty and spatial variability in estimation of long-term sulfate dynamics in a region. Water
Resour. Res. 22:1293-1302.
Hornberger, G.J., B.J. Cosby, and R.F. Wright. 1989. Historical reconstructions and future forecasts of
regional surface water acidification in southernmost Norway. Water Resour. Res. 25:2009-2018.
341
-------
Hosker, R.P., and J.D. Womack. 1986. Simple meteorological and chemical filterpack monitoring system
for estimating dry deposition of gaseous pollutants, pp. 23-29. In: Proceedings of the Fifth Annual
National Symposium on Recent Advances in Pollutant Monitoring of Ambient Air and Stationary
Sources, January. EPA/600/9-85/029. Atmospheric Turbulence and Diffusion Division,
NOAA/ERL/Air Resources Laboratory, Oak Ridge, TN.
Huang, P.M. 1977. Feldspars, olivines, pyroxenes and amphiboles, pp. 553-602. In: J.B. Dixon and S.B.
Weed, eds., Minerals in Soil Environments. Soil Science Society of America, Madison, Wl.
Hunsaker, C.T., S.W. Christensen, JJ. Beauchamp, R.J. Olson, R.S. Turner, and J.L. Malanchuk. 1986.
Empirical Relationships Between Watershed Attributes and Headwater Lake Chemistry in the
Adirondack Region. Environ. Sci. Div. Pub. No. 2884. ORNL/TM-9838. Oak Ridge National
Laboratory, Oak Ridge, TN. 123 pp.
Hutchlnson, T.C., and M. Havas. 1986. Recovery of previously acidified lakes near Coniston, Canada,
following reductions in atmospheric sulphur and metal emissions. Water Air Soil Pollut. 23:319-333.
Hynes, H.B.N. 1975. Edgardo Baldi memorial lecture: The stream and its valley. Verh. Internal. Verein.
Umnol. 19:1.
Jenkins, A., B.J. Cosby, R.C. Ferrier, TAB. Walker, and J.D. Miller. 1990a. Modelling stream acidification
in afforested catchments: An assessment of the relative effects of acid deposition and
afforestation. J. Hydrol. 120:163-181.
Jenkins, A., P.G. Whitehead, B.J. Cosby, and H.J.B. Birks. 1990b. Modelling long term acidification: A
comparison with diatom reconstructions and the implications for reversibility. Phil. Trans. R. Soc.
Lond. B. 327:435-440.
Jenkins, A., P.G. Whitehead, T.J. Musgrove, and B.J. Cosby. 1990c. A regional model of acidification in
Wales. J. Hydrol. 116:403-416.
Jenne, E.A., LE. Eary, LW. Vail, D.C. Gin/in, A.M. Liebetrau, LF. Hibler, T.B. Miley, and M.J. Monsour.
1989. An Evaluation and Analysis of Three Dynamic Watershed Acidification Codes (MAGIC, ETD,
and ILWAS). PNL-6687/UC-11. Richland, WA.
Johnson, D.W., and D.W. Cole. 1980. Anion mobility in soils: Relevance to nutrient transport from forest
ecosystems. Environ. Internal. 3:79-90.
Johnson, D.W., and J.O. Reuss. 1984. Soil mediated effects of atmospherically deposited sulphur and
nitrogen. Phil. Trans. R. Soc. Lond. 3056:383-392.
Johnson, D.W., and D.E. Todd. 1983. Relationships among iron, aluminum, carbon, and sulfate in a
variety of forest soils. Soil Sci. Soc. Am. J. 47:792-800.
342
-------
Johnson, D.W., and D.E. Todd. 1987. Nutrient export by leaching and whole-tree harvesting in a loblolly
pine and mixed oak forest. Plant and Soil 102:99-109.
Johnson, D.W., J.W. Hornbeck, J.M. Kelly, W.T. Swank, and D.E. Todd. 1980. Regional patterns of soil
sulfate accumulation: Relevance to ecosystem sulfur budgets, pp. 507-519. In: D.S. Shriner, C.R.
Richmond and S.E. Lindberg, eds. Atmospheric Sulfur Deposition: Environmental Impact and
Health Effects. Ann Arbor Science, Ann Arbor, Ml.
Johnson, D.W., J. Turner, and J.M. Kelly. 1982. The effects of acid rain on forest nutrient status. Water
Resour. Res. 18:449-461.
Johnson, D.W., G.S. Henderson, and D.E. Todd. 1988. Changes in nutrient distribution in forests and
soils of Walker Branch watershed, Tennessee, over an eleven-year period. Biogeochemistry
5:275-293.
Johnson, M.G., P.W. Shaffer, D.L Stevens, K.W. Thornton, and R.S. Turner. 1990. Predicting and
Forecasting Surface Water Acidification: A Plan to Assess Data Aggregation Effects. U.S.
Environmental Protection Agency, Environ. Res. Lab., Corvallis, OR. 93 pp.
Johnson, R.A., and D.W. Wichern. 1982. Applied Multivariate Statistical Analysis. Prentice-Hall, Inc.,
Englewood Cliffs, NJ. 594 pp.
Kanciruk, P., J.M. Eilers, R.A. McCord, D.H. Landers, D.F. Brakke, and R.A. Linthurst. 1986.
Characteristics of Lakes in the Eastern United States, Volume III: Data Compendium of Site
Characteristics and Chemical Variables. EPA/600/4-86/007c. U.S. Environmental Protection
Agency, Washington, DC. 439 pp.
Kaufmann, P.R., AT. Herlihy, J.W. Elwood, M.E. Mitch, W.S. Overton, M.J. Sale, J.J. Messer, K.A.
Cougar, D.V. Peck, K.H. Reckhow, A.J. Kinney, S.J. Christie, D.D. Brown, C.A. Hagley, and H.I.
Jager. 1988. Chemical Characteristics of Streams in the Mid-Atlantic and Southeastern United
States, Volume I: Population Descriptions and Physico-Chemical Relationships.
EPA/600/3-88/021 a. U.S. Environmental Protection Agency, Washington, DC. 397 pp.
Keller, W., and J.R. Pitblado. 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.
Kelly, C.A., J.W.M. Rudd, R.H. Hesslein, D.W. Schindler, P.J. Dillon, C.T. Driscoll, S.A. Gherini, and R.E.
Hecky. 1987. Prediction of biological acid neutralization in acid-sensitive lakes. Biogeochemistry
3:129-140.
Kern, J., ManTech Environmental Technology, Inc., Corvallis, Oregon, personal communication.
343
-------
Kern, J.S., and J.J. Lee. 1990. DDRP: Field Operations and Quality Assurance Report for Soil Sampling
in the Mid-Appalachian Region of the US. EPA/600/3-90/045. US Environmental Protection
Agency, Washington, DC. 56 pp
Kern, J.S., D.A. Lammers, D.L Cassell, and W.G. Campbell. 1990. DDRP: Field Operations and Quality
Assurance Report for Watershed Mapping in the Mid-Appalachian Region of the US. EPA/600/3-
90/036. U.S. Environmental Protection Agency, Washington, DC. 169 pp.
Khanna, P.K., and F. Beese. 1978. The behavior of sulfate salt input in podzolic brown earth. Soil Sci.
125:16-22.
Khanna, P.K., J. Prenzel, K.J. Meiwes, B. Ulrich, and E. Matzner. 1987. Dynamics of sulfate retention by
acid forest soils in an acidic deposition environment. Soil Sci. Soc. Am. J. 51:446-452.
Kittrick, J.A. 1977. Mineral equilibria and the soil system, pp. 1-26. In: J.B. Dixon and S.B. Weed, eds.,
Minerals in Soil Environments. Soil Science Society of America, Madison, Wl.
Krug, E.G. 1989. Assessment of the theory and hypotheses of the acidification of watersheds. Illinois
State Water Survey Division, SWS Contract Report 457. 25? pp.
Krug, E.C., and C.R. Frink. 1983. Acid rain on acid soil: A new perspective. Science 221:520-525.
Krug, E.C., and P.J. Isaacson. 1984. Comparison of water and dilute acid treatment on organic and
inorganic chemistry of leachate from organic-rich horizons of an acid forest soil. Soil Sci.
137:370-378.
Krug, E.C., P.J. Isaacson, and C.R. Frink. 1985. Appraisal of some current hypotheses describing
acidification of watersheds. J. Air Pollut. Contr. Assoc. 35:109-114.
Krug, W.R., W.A. Gebert, D.J. Graczyk, D.L. Stevens, B.P. Rochelle, and M.R. Church. 1990. Map of
Mean Annual Runoff for the Northeastern, Southeastern, and Mid-Atlantic United States, Water
Years 1951-80. U.S. Geological Survey Water Resour. Invest. Report No. 88-4094, Madison, Wl.
11pp.
Kulp, L 1987. Interim assessment, the causes and effects of acidic deposition. Volume I: Executive
Summary. National Acid Precipitation Assessment Program, Washington, DC.
Lammers, DA, D.L Stevens, and M.G. Johnson. 1987. Soil mapping data management system. In:
Agronomy Abstracts, American Society of Agronomy, Atlanta, November 29 - December 4, 1987.
Lammers, D.A., J.J. Lee, and J.A. Ferwerda. 1988. Mapping Protocols for Direct/Delayed Response
Project Soil Mapping Activities in the Mid-Appalachian Region. U.S. Environmental Protection
Agency, Corvallis, OR. 39 pp.
344
-------
Landers, D.H., J.M. Eilers, D.F. Brakke, W.S. Overton, P.E. Kellar, M.E. Silverstein, R.D. Schonbrod, R.E.
Crowe, R.A. Linthurst, J.M. Omernik, S.A. league, and E.P. Meier. 1987. Characteristics of Lakes in
the Western United States, Volume I: Population Descriptions and Physico-Chemical Relationships.
EPA/600/3-86/054a. U.S. Environmental Protection Agency, Washington, DC. 176 pp.
Lasaga, A.C. 1981. Rate laws of chemical reactions, pp. 1-68. In: A.C. Lasaga and R.J. Kirkpatrick, eds.,
Kinetics of Geochemical Processes. Reviews in Mineralogy Series, Volume 8. Mineralogical
Society of America, Washington, D.C.
Lau, W.M., and S.J. Mainwaring. 1985. The determination of soil sensitivity to acid deposition. Water, Air,
Soil Pollut. 25:451-464.
LaZerte, B.D., and P.J. Dillon. 1984. Relative importance of anthropogenic versus natural sources of
acidity in lakes and streams of central Ontario. Can. J. Fish. Aquat. Sci. 41:1664-1677.
Lee, J.J., and D.E. Weber. 1983. Effects of sulfuric acid rain on decomposition rate and chemical
element content of hardwood leaf litter. Can. J. Bot. 61:872-879.
Lee, J.J., D.A. Lammers, M.G. Johnson, M.R. Church, D.L Stevens, D.S. Coffey, R.S. Turner, LJ. Blume,
LH. Liegel, and G.R. Holdren. 1989a. Watershed surveys to support an assessment of the regional
effect of acidic deposition on surface water chemistry. Environ. Manage. 13:95-108.
Lee, J.J., D.A. Lammers, D.L. Stevens, K.W. Thornton, and K.A. Wheeler. 1989b. Classifying soils for
acidic deposition aquatic effects: A scheme for the northeastern U.S. Soil Sci. Soc. Am. J.
53:1153-1163.
Lee, J.J., D.L. Stevens, J.S. Kern, and D.A. Lammers. 1990. Direct/Delayed Response Project: Mid-
Appalachian Soil Correlation and Sampling Class Development Workshop. EPA/600/3-90/037. US
Environmental Protection Agency, Washington, DC. 43 pp.
Lee, S. 1987. Uncertainty Analysis for Long-Term Acidification of Lakes in Northeastern USA. Ph.D.
Thesis. University of Iowa, Iowa City.
Lee, S., and J.L. Schnoor. 1988. Reactions that modify chemistry in lakes of the NSWS. Environ. Sci.
Technol. 22:190-197.
Lepisto, A., P.G. Whitehead, C. Neal, and B.J. Cosby. 1988. Modeling effects of acid deposition:
Estimation of long-term water quality responses in forested catchments in Finland. Nordic
Hydrology. 19:99-120.
Liegel, L.H., D. Cassell, R. Church, D. Stevens, and P. Shaffer. 1991. Regional characteristics of land
use in the Northeast and Southern Blue Ridge Province: Associations with acid rain effects on
surface-water chemistry. Environ. Manage. 15:269-279.
345
-------
Likens, G.E., F.H. Bormann, R.S. Pierce, J.S. Eaton, and N.M. Johnson. 1977. Biogeochemistry of a
Forested Ecosystem. Springer-Verlag, Inc., New York, NY. 146 pp.
Llndberg, S.E., G.M. Lovett, D.D. Richter, and D.W. Johnson. 1986. Atmospheric deposition and canopy
interactions of major ions in a forest. Science 231:141-145.
Llnthurst, R.A., D.H. Landers, J.M. Eilers, D.F. Brakke, W.S. Overton, E.P. Meier, and R.E. Crowe. 1986.
Characteristics of Lakes in the Eastern United States, Volume I: Population Descriptions and
Physico-Chemical Relationships. EPA/600/4-86/007a. U.S. Environmental Protection Agency,
Washington, DC: 136 pp.
Lynch, D.D., and N.B. Dise. 1985. Sensitivity of Stream Basins in Shenandoah National Park to Acid
Deposition. USGS Water Resources Investigation Report 85-4115. U.S. Geological Survey,
Richmond, VA. 61 pp.
Mast, M.A., and J.I. Drever. 1987. The effect of oxalate on the dissolution rate of oligoclase and tremolite.
Geochim. Cosmochim. Acta 51:2559-2568.
Matalas, N.C., and B. Jacobs. 1964. A Correlation Procedure for Augmenting Hydrology Data. U.S.
Geological Survey Professional Paper 434, Washington, DC. 7 pp.
McColl, J.G., and A.A. Pohlman. 1986. Soluble organic acids and their chelating influence on Al and
other metal dissolution from forest soils. Water Air Soil Pollut. 31:917-927.
Messer, J.J., C.W. Ariss, J.R. Baker, S.K. Drous6, K.N. Eshleman, P.R. Kaufmann, R.A. Linthurst, J.M.
Omernik, W.S. Overton, M.J. Sale, R.D. Schonbrod, S.M. Stambaugh, and J.R. Tuschall, Jr. 1986.
National Stream Survey Phase I, Pilot Survey. EPA/600/4-86/026. U.S. Environmental Protection
Agency, Washington, DC. 179 pp.
Mitchell, M.J., M.B. David, D.G. Maynard, and S.A. Telang. 1986. Sulfur constituents in soils and streams
of a watershed in the Rocky Mountains of Alberta. Can. J. For. Res. 16:315-320.
Mohnen, V.A. 1988. The challenge of acid rain. Scientific American 259:30-38.
Morgan, M.D. 1990. Streams in the New Jersey Pinelands directly reflect changes in atmospheric
deposition chemistry. J. Environ. Qual. 19:296-302.
Mortenson, D.C. 1989a. Geographic Information System Documentation of Watershed Data for
Direct/Delayed Response Project - Northeast Database. EPA/600/3-89/001 and Project Summary.
U.S. Environmental Protection Agency, Corvallis, OR. 35 pp.
Mortenson, D.C. 1989b. Geographic Information System Documentation of Watershed Data for
Direct/Delayed Response Project - Southern Blue Ridge Province Database. EPA/600/3-89/002
and Project Summary. U.S. Environmental Protection Agency, Corvallis, OR. 25 pp.
346
-------
Munson, R., Tetra-Tech, Inc., Hadley, Massachusetts, personal communication.
Musgrove, T.J., P.G. Whitehead, B.J. Cosby, G.M. Hornberger, and R.F. Wright. 1990. Regional modeling
catchments in the Galloway region in southwest Scotland. In: J. Kamari, ed. Impact Models to
Assess Regional Acidification. Kluwer Academic Publishers, Amsterdam.
NAS (National Academy of Sciences). 1984. Acid deposition: Processes of lake acidification. Summary of
a Discussion. National Research Council Commission on Physical Sciences, Mathematics, and
Resources. Environmental Studies Board, Panel on Processes of Lake Acidification. National
Academy Press, Washington, DC. 1 1 pp.
NAS (National Academy of Sciences). 1986. Acid Deposition: Long Term Trends. National Research
Council Commission on Physical Sciences, Mathematics, and Resources. Environmental Studies
Board. National Academy Press. Washington, D.C.
Neal, C., P.G. Whitehead, R. Neale, and B.J. Cosby. 1986. Modeling the effect of acid deposition and
conifer afforestation on stream acidity in the British uplands. J. Hydrol. 86:15-26.
Neary, B.P., and P.J. Dillon. 1988. Effects of sulphur deposition on lake-water chemistry in Ontario,
Canada. Nature 333:340-343.
Neller, J.R. 1959. Extractable sulfate - sulfur in soils of Florida in relation to amount of clay in the profile.
Soil Sci. Soc. Am. J. 23:346-348.
Nevell, N., and M. Wainwright. 1986. Microbial growth on polluted leaf and litter leachates: Effects on
leachate pH and sulphate concentration. Plant and Soil. 92:303-308.
Nikolaidis, N.P., H. Rajaram, J.L Schnoor, and K.P. Georgakakos. 1988. A generalized soft water
acidification model. Water Resour. Res. 24:1983-1996.
Nodvin, S.C., C.T. Driscoll, and G.E. Likens. 1986. The effect of pH on sulfate adsorption by a forest soil.
Soil Sci. 142:69-75.
Nordstrom, O.K. 1982. The effects of sulfate on aluminum concentrations in natural waters: Some
stability relations in the system AI^ - SQj - h^O at 298 K. Geochim. Cosmochim. Acta
46:681-692.
Norton, S.A., R.F. Wright, J.S. Kahl, and J.P. Scofield. The MAGIC simulation of surface water
acidification at, and first results from, the Bear Brook Watershed Manipulation, Maine, USA.
Environ. Poll. In press.
NRC (National Research Council). 1988. Estimating probabilities of extreme floods. Methods and
Recommended Research. National Academy Press. Washington, DC.
347
-------
Nyborg, M. 1978. Sulfur pollution and soil, pp. 359-390. In: J.O. Nriagu, ed. Sulfur in the Environment,
Part II: Ecological Impacts. Wiley-lnterscience, New York, NY.
Nye, P.H., and D.J. Greenland. 1960. The Soil Under Shifting Cultivation. Commonwealth Bureau of Soils
Tech. Comm. No. 51. Commonwealth Agricultural Bureaux, Farnham Royal, Bucks.
Olsen, A., U.S. EPA, Corvallis, Oregon, personal communication.
Olsen, A.R., E.G. Voldner, D.S. Bigelow, W.H. Chan, T.L. Clark, M.A. Lusis, P.K. Misra, and R.J. Vet.
1990. Unified wet deposition data summaries for North America: Data summary procedures and
results for 1980-1986. Atmos. Environ. 24A:661-672.
Overton, W.S. 1987. A Sampling and Analysis Plan for Streams in the National Surface Water Survey.
Technical Report 117. Department of Statistics, Oregon State University, Corvallis, Oregon, 97331.
50 pp.
Paces, T. 1973. Steady-state kinetics and equilibrium between ground water and granitic rock. Geochim.
Cosmochim. Acta 37:2641-2663.
Papp, M.L, and R.D. Van Remortel. 1990. Direct/Delayed Response Project: Preparation Laboratory
Operations and Quality Assurance Report for the Mid-Appalachian Soil Survey. EPA/600. US
Environmental Protection Agency, Washington, DC. 226 pp.
Papp, M.L, R.D. van Remortel, CJ. Palmer, G.E. Byers, B.A. Schumacher, R.L Slagle, J.V. Burton, K.C.
Shines, J.E. Teberg and M.J. Miah. 1989. Direct/Delayed Response Project: Quality Assurance
Plan for Preparation and Analysis of Soils from the Mid-Appalachian Region of the United States.
EPA/600/4-89/031. U.S. Environmental Protection Agency, Washington, D.C. 226 pp.
Parfitt, R.L, and R.S.C. Smart. 1978. The mechanism of sulfate adsorption on iron oxides. Soil Sci. Soc.
Am. Proa 42:48-50.
Peters, N.E., and C.T. Driscoll. 1987. Hydrogeologic controls of surface-water chemistry in the
Adirondack Region of New York State. Biogeochemistry 3:163-180.
Rajan, S.S.S. 1978. Sulfate adsorbed on hydrous alumina, ligands displaced, and changes in surface
charge. Soil Sci. Soc. Am. J. 42:39-44.
Rajan, S.S.S. 1979. Adsorption and desorption of sulfate and charge relationships in allophanic clays.
Soil Sci. Soc. Am. J. 43:65-69.
Reuss, J.O. 1983. Implications of the calcium-aluminum exchange system for the effect of acid
precipitation in soils. J. Environ. Qual. 14:26-31.
348
-------
Reuss, J.O., and D.W. Johnson. 1985. Effect of soil processes on the acidification of water by acid
deposition. J. Environ. Qual. 14:26-31.
Reuss, J.O., and D.W. Johnson. 1986. Acid Deposition and the Acidification of Soils and Waters.
Ecological Studies Volume 59. Springer-Verlag, Inc., New York, NY.
Richter, D.D., P.J. Comer, K.S. King, H.S. Sawin, and D.S. Wright. 1988. Effects of low ionic strength
solutions on pH of acid forested soils. Soil Sci. Soc. Am. J. 52:261-264.
Rochelle, B.R., and M.R. Church. 1987. Regional patterns of sulfur retention in watersheds of the eastern
U.S. Water Air Soil Pollut. 36:61-73.
Rochelle, B.P., M.R. Church, and M.B. David. 1987. Sulfur retention at intensively studied sites in the U.S.
and Canada. Water Air Soil Pollut. 33:73-84.
Rochelle, B.P., C.I. Lift, W.G. Campbell, D.L Cassell, M.R. Church, and R.A. Nusz. 1989a. Relating map
geomorphic/hydrologic parameters to surface water chemistry regionally: Relative to acidic
deposition. J. Hydro!. 112:103-120.
Rochelle, B.P., D.L. Stevens, and M.R. Church. 1989b. Uncertainty analysis of runoff estimates from a
runoff contour map. Water Resour. Bull. 25:491 -498.
Rosenbrock, H.H. 1960. An automatic method for finding the greatest or least value of a function.
Comput. J. 3:175-184.
Rudd, J.W.M., C.A. Kelly, and A. Furutani. 1986a. The role of sulfate reduction in long term accumulation
of organic and inorganic sulfur in lake sediments. Limnol. Oceanogr. 36:1281-1291.
Rudd, J.W.M., C.A. Kelly, V. St. Louis, R.H. Hesslein, A. Furutani, and M.H. Holoka. 1986b. Microbial
consumption of nitric and sulfuric acids in north temperate-lakes. Limnol. Oceanogr. 31:1267-1280.
Ryan, P.F., G.M. Hornberger, B.J. Cosby, J.N Galloway, J.R. Webb, and E.B. Rastetter. 1989. Changes
in the chemical composition of stream water in two catchments in the Shenandoah National Park,
Virginia, in response to atmospheric deposition of sulfur. Wat. Resour. Res. 25:2091 -2099.
Sale, M.J., P.R. Kaufmann, H.I. Jager, J.M. Coe, K.A. Cougan, A.J. Kinney, M.E. Mitch, and W.S.
Overton. 1988. Chemical Characteristics of Streams in the Mid-Atlantic and Southeastern United
States, Volume II: Streams Sampled, Descriptive Statistics, and Compendium of Physical and
Chemical Data. EPA/600/3-88/021 b. U.S. Environmental Protection Agency, Washington, DC. 597
pp.
Sanders, F.E., and P.B.H. Tinker. 1975. Adsorption of sulphate by a sandy loam soil (calcic cambisol).
Geoderma 13:317-324.
349
-------
SAS Institute, Inc. 1987. SAS/STAT Guide for Personal Computers, Version 6 Edition. SAS Institute, Inc.,
Gary, NC. 1028 pp.
SAS Institute, Inc. 1988. SAS/STAT User's Guide, Release to 6.03 Edition. SAS Institute, Inc., Gary, NC.
Schindler, D.W. 1986. The significance of in-lake production of alkalinity. Water Air Soil Pollut.
30:931-946.
Schindler, D.W. 1988. The effects of acid rain on freshwater ecosystems. Science 239:149-157.
Schindler, S.C., M.J. Mitchell, T.J. Scott, R.D. Fuller, and C.T. Driscoll. 1986a. Incorporation of 35S-sulfate
into inorganic and organic constituents of two forest soils. Soil Sci. Soc. Am. J. 40:457-462.
Schindler, D.W., M.A. Turner, M.P. Stainton, and G.A. Linsey. 1986b. Natural sources of acid neutralizing
capacity fn low-alkalinity lakes of the Precambrian Shield. Science 232:844-847.
Schnoor, J.L, 8. Lee, and N.P. Nikolaidis. 1986. Lake resources at risk to acidic deposition in the
eastern United States. Water Air Soil Pollut. 31:1091-1101.
Schofield, C.L 1976. Acidification of Adirondack Lakes by Atmospheric Precipitation: Extent and
Magnitude of the Problem. Final report, Project F-28-R. New York State Dept. Environ. Conserv.,
Albany, NY.
Sellkop, S., AnaJytical Sciences, Inc., Durham, North Carolina, personal communication.
Selp, H.M. 1980. Acidification of freshwaters - sources and mechanisms, pp. 358-366. In: D. Drabl0s and
A. Tollan, eds. Ecological Impact of Acid Precipitation: Proceedings of an International Conference,
Sandefjord, March 11-14. SNSF Project, Oslo-As, Norway.
Shaffer, P.W. 1989. Organic blocking of sulfate adsorption by a Virginia forest soil. Agronomy Abstracts.
p. 311.
Shaffer, P.W., and M.R. Church. 1989. Terrestrial and in-lake contributions to the alkalinity budgets of
drainage lakes: An assessment of regional differences. Can. J. Fish. Aquat. Sci. 46:509-515.
Shaffer, P.W., and D.L Stevens, Jr. 1991. Procedures for aggregation of nonlinear data: an assessment
using soil sulfate adsorption isotherms. Soil Science Society of America Journal. 55:684-692.
Shaffer, P.W., R.P. Hooper, K.N. Eshleman, and M.R. Church. 1988. Watershed vs. in-lake alkalinity
generation: A comparison of rates using input-output studies. Water, Air, Soil Pollut. 39:263-273.
Simons, T.J., and D.C.L Lam. 1980. Some limitations of water quality models for large lakes: A case
study of lake Ontario. Water Resour. Res. 16:105-116.
350
-------
Smith, R.A., and R.B. Alexander. 1983. Evidence for Acid-Precipitation Induced Trends in Stream
Chemistry at Hydrologic Bench-Mark Stations. Geological Survey Circular 910. U.S. Geological
Survey, Alexandria Virginia, 12 pp.
Soil Survey Staff. 1951. Soil Survey Manual. USDA Soil Conservation Service, Agricultural Handbook No.
18. U.S. Gov. Print. Office, Washington, DC.
Soil Survey Staff. 1975. Soil Taxonomy: A Basic System of Soil Classification for Making and Interpreting
Soil Surveys. USDA Soil Conservation Service, Agricultural Handbook No. 436. U.S. Gov. Print.
Office, Washington, DC. 754 pp.
Soil Survey Staff. 1983. National Soils Handbook. USDA Soil Conservation Service, Agricultural
Handbook No. 430. U.S. Gov. Print. Office, Washington, DC.
Soil Survey Staff. 1984. Procedures for Collecting Soil Samples and Methods of Analysis for Soil Survey.
USDA Soil Conservation Service, SSIR 1. U.S. Gov. Print. Office, Washington, DC. 68 pp.
Solomon, K.P., and T.E. Cerling. 1987. The annual carbon dioxide cycle in a montane soil: Observations,
modeling, and implications for weathering. Water Resour. Res. 25:2257-2265.
Stensland, G., Illinois State Water Survey, Champaign, Illinois, personal communication.
Stevenson, F.J. 1982. Humus Chemistry: Genesis, Composition, Reactions. Wiley-lnterscience, New
York. NY. 443 pp.
Strickland, T., ManTech Environmental Technology, Inc., Corvallis, Oregon, personal communication.
Strickland, T.C., and J.W. Fitzgerald. 1984. Formation and mineralization of organic sulfur in forest soils.
Biogeochemistry 1:79-95.
Strickland, T.C., J.W. Fitzgerald, J.T. Ash, and W.T. Swank. 1987. Organic sulfur transformations and
sulfur pool sizes in soil and litter from a Southern Appalachian hardwood forest. Soil Sci.
143:453-458.
Stuanes, A.O. 1984. A simple extraction as an indicator of a soil's sensitivity to acid precipitation. Acta
Agric. Scand. 34:113-127.
Sullivan, T.J., J.M. Eilers, M.R. Church, D.J. Blick, K.N. Eshleman, D.H. Landers, and M.S. DeHaan. 1988.
Atmospheric wet sulphate deposition and lakewater chemistry. Nature 331:607-609.
Sullivan, T.J., J.A. Bernert, E.A. Jenne, J.M. Eilers, B.J. Cosby, D.F. Charles, and A.R. Selle. 1991.
Comparison of MAGIC and DIATOM Paleolimnological Model Hindcasts of Lakewater Acidification
in the Adirondack Region of New York. PNL-7487, Pacific Northwest Laboratory, Richland,
Washington.
351
-------
Swank, W.T., and D.A. Crossley, Jr., eds. 1988. Ecological Studies, Volume 66: Forest Hydrology and
Ecology at Coweeta. Springer-Verlag, New York, NY. 469 pp.
Swank, W.T., and J.B. Waide. 1988. Characterization of baseline precipitation and stream chemistry and
nutrient budgets for control watersheds, pp. 57-79. In: W.T. Swank and D.A. Crossley, Jr., eds.
Ecological Studies, Volume 66: Forest Hydrology and Ecology at Coweeta. Springer-Verlag, New
York, NY.
Swank, W.T., J.W. Fitzgerald, and J.T. Ash. 1984. Microbial transformations of sulfate in forest soils.
Science 223:182-184.
Tamm, C.O., and L Hallbacken. 1988. Changes in soil acidity in two forest areas with different acid
deposition: 1920s to 1980s. Ambio 17:56-61.
Tang, A.J.S., W.H. Chan, D.H.S. Chung, and M.A. Lusis. 1986. The Acid Precipitation in Ontario Study:
Precipitation and Air Concentration and Wet and Dry Deposition Fields of Pollutants in Ontario,
1983. Report #ARB-008-86-AQM, Ontario Ministry of the Environment.
Thornton, K., D. Marmorek, P. Ryan, K. Heltcher, and D. Robinson. 1990. Methods for Projecting Future
Changes in Surface Water Acid-base Chemistry. SOS/T Report 14. National Acid Precipitation
Assessment Program
Troedsson, T. 1980. Long-term changes of forest soils. Ann. Agr. Fenniae. 19:81-84.
Turner, R.S., R.B. Cook, H. Van Miegroet, D.W. Johnson, J.W. Elwood, O.P. Bricker, S.E. Lindberg, and
G.M. Hornberger. 1990. Watershed and Lake Processes Affecting Chronic Surface Water Acid-
Base Chemistry, NAPAP SOS/T Report 10, In: Acidic Deposition: State of Science and"
Technology. National Acid Precipitation Assessment Program, Volume II, Washington, DC.
Ulrich, B. 1983. Soil acidity and its relation to acid deposition. In: B. Ulrich, and J. Pankrath, eds. Effects
of Accumulation of Air Pollutants in Forest Ecosystems. D. Reidel Publ. Co., Boston, MA.
U.S. Environmental Protection Agency, Office of Policy, Planning and Evaluation, Washington, D.C.,
personal communication.
U.S. Environmental Protection Agency. 1984. National Surface Water Survey: National Stream Survey;
Draft 1985 Research Plan. Environmental Research Laboratory, Corvallis, Oregon. 134 pp.
USEPA (U.S. Environmental Protection Agency). 1988. Statements of Work for Soil Analysis in the
Direct/Delayed Response Project Mid-Appalachian Soil Survey. Attachment A in Invitations for Bid,
Solicitations W802498D1 and W802621D1. U.S. Environmental Protection Agency, Las Vegas, NV.
313 pp. and 36 pp.
352
-------
U.S. Geological Survey. 1969. Water Resources Data, Virginia, Water Year 1968. U.S. Geological Survey
Water Report VA-68-1.
U.S. Geological Survey. 1970. Water Resources Data, Virginia, Water Year 1969. U.S. Geological Survey
Water Report VA-69-1.
VanLoon, G.W. 1984. Acid rain and soil. Can. J. Physiol. Pharmacol. 62:991-997.
Vanselow, A.P. 1932. Equilibria of the base-exchange reactions of bentonites, permutites, soil colloids,
and zeolites. Soil sci. 33:95-113.
Velbel, M.A. 1985. Geochemical mass balances and weathering rates in forested watersheds of the
Southern Blue Ridge. Am.J. Sci. 285:904-930.
Velbel, M.A. 1986a. The mathematical basis for determining rates of geochemical and geomorphic
processes in small forested watersheds by mass balance: Examples and implications, pp. 439-452.
In: S.M. Colman and D.P. Dethier, eds. Rates of Chemical Weathering of Rocks and Minerals.
Academic Press, Orlando, FL.
Velbel, M.A. 1986b. Influence of surface area, surface characteristics, and solution composition of
feldspar weathering rates, pp. 615-634. In: J.A. Davis and K.F. Hayes, eds. Geochemical Processes
at Mineral Surfaces. American Chemical Society Symposium Series No. 323, Washington, DC.
Vong, R., S. Cline, G. Reams, J. Bernert, D. Charles, J. Gibson, T. Haas, J. Moore, R. Husar, A. Olsen, J.
Simpson, and S. Seilkop. 1989. Regional Analysis of Wet Deposition for Effects Research.
EPA/600/3-89/030. U.S. Environmental Protection Agency, Environmental Research Laboratory,
Corvallis, OR. 57 pp.
Waide, J., FTN & Associates, Little Rock, Arkansas, personal communication.
Waide, J.B., and W.T. Swank. 1987. Patterns and Tends in Precipitation and Stream Chemistry at the
Coweeta Hydrologic Laboratory. Aquatic Effects Task Group VI Peer Review. Volume II. May 17-
23, New Orleans, LA.
Wampler, S.J., and A.R. Olsen. 1987. Spatial estimation of annual wet acid deposition using
supplemental precipitation data, pp. 248-251. In: Tenth Conference on Probability and Statistics,
Oct. 4-5, Edmonton, Atla., Canada. American Meteorological Society, Boston, MA.
Weaver, G.T., P.K. Khanna, and F. Beese. 1985. Retention and transport of sulfate in slightly acid forest
soil. Soil Sci. Soc. Am. J. 49:746-750.
Webb, J.R., B.J. Cosby, J.N. Galloway, and G.M. Hornberger. 1989. Acidification of native brook trout
streams in Virginia. Wat. Resour. Res. 25:1367-1377.
353
-------
Welder, R.K., and G.E. Lang. 1988. Cycling of inorganic and organic sulfur in peat from Big Run Bog,
West Virginia. Biogeochemistry 5:221-242.
Whitehead, P.G., S. Bird, M. Hornung, J. Cosby, C. Neal, and P. Paricos. 1988a. Stream acidification
trends in the Welsh uplands - A modelling study of the Uyn Brianne catchments. J. Hydrol.
101:191-212.
Whitehead, P.G., B. Reynolds, M. Hornung, C. Neal, J. Cosby, and P. Paricos. 1988b. Modelling long
term stream acidification trends in Upland Wales at Plynlimon. Hydrol. Proc. 2:357-368.
Wigington, P.J., Jr., T.D. Davies, M. Tranter, and K.N. Eshleman. 1990. Episodic Acidification of Surface
Waters Due to Acidic Deposition. NAPAP SOS/T Report 12. In: Acidic Deposition: State of
Science and Technology. National Acid Precipitation Assessment Program, Volume II, Washington,
DC.
Wollast, R. 1967. Kinetics of the alteration of K-feldspar in buffered solutions at low temperature.
Geochim. Cosmochim. Acta 31:635-648.
Wright, R.F., B.J. Cosby, G.M. Hornberger, and J.N. Galloway. 1986. Comparison of paleolimnological
with MAGIC model reconstructions of water acidification. Water Air Soil Pollut. 30:367-380.
Wright, R.F., B.J. Cosby, M.B. Platen, and J.O. Reuss. 1990. Evaluation of an acidification model with
data from manipulated catchments in Norway. Nature. 343:53-55.
Wright, R.F., B.J. Cosby, and G.M. Hornberger. 1991. A regional model of lake acidification in
southernmost Norway. Ambio. 20:222-225.
Wright, R.F., E. Lotse, and A. Semb. 1988. Reversibility of acidification shown by whole-catchment
experiments. Nature 334:670-675.
354
-------
SECTION 13
GLOSSARY
13.1 ABBREVIATIONS AND SYMBOLS
13.1.1 Abbreviations
ADS
AERP
ANC
APIOS
AREAL-RTP
BMD
BMK
END
BNS
BXP
BXW
CDF
CFP
Cl
CMS
DDRP
DOC
DQO
ELS-I
EMSL-LV
EPA
EPRI
ERL-C
ERP
ETD
FLW
GIS
Acid Deposition System
Aquatic Effects Research Program
Acid neutralizing capacity
Acid Precipitation in Ontario Study conducted by Ontario Ministry of the
Environment
USEPA Atmospheric Research and Exposure Assessment Laboratory -
Research Triangle Park
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
Cumulative distribution function
A DDRP soil sampling class in the M-APP Region
Confidence interval
A DDRP soil sampling class in the M-APP Region
Direct/Delayed Response Project
Dissolved organic carbon
Data quality objective
Eastern Lake Survey-Phase I
USEPA Environmental Monitoring and Systems Laboratory - Las Vegas
U.S. Environmental Protection Agency
Electric Power Research Institute .
USEPA Environmental Research Laboratory - Corvallis
Episodic Response Project
Enhanced Trickle Down Model
A DDRP soil sampling class in the M-APP Region
Geographic Information System
355
-------
HST
A DDRP soil sampling class in the M-APP Region
ILWAS
IQR
LA!
LEVIS
LTA
MAGIC
MLRA
MQO
NAS
NADP/NTN
NAPAP
NCDC
NE
NOAA
NRC
NSS-I
NSWS
ORNL
PE
PNL
- Integrated Lake/Watershed Acidification Study
- Interquartile range
— Leaf area index
- Laboratory Entry and Verification System
- Long-term annual average deposition
- Model for Acidification of Groundwater in Catchments
— Major land resource areas
- Measurement quality objective
— National Academy of Sciences
- National Acid Deposition Program/National Trends Network
— National Acid Precipitation Assessment Program
- National Climatic Data Center
— Northeast Region
- National Oceanographic and Atmospheric Administration
- National Research Council
— National Stream Survey-Phase I
- National Surface Water Survey
- Oak Ridge National Laboratory, Tennessee
- Performance evaluation
- Battelle-Pacific Northwest Laboratories
QA
QC
RADM
RELMAP
RCC
RMSE
RSD
Quality assurance
Quality control
Regional Acid Deposition Model
Regional Lagrangian Model of Air Pollution
Regional Coordinator/Correlator
Root mean square error
Relative standard deviation
SAS
SBRP
SCS
SE
SOBC
Statistical Analysis System
Southern Blue Ridge Province
Soil Conservation Service
Standard error
Sum of base cations
356
-------
TMY
TPD
TWO
TWM
TXP
TXW
TY
UMW
USDA
USFS
USGS
UTM
Typical meteorological year
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
A DDRP soil sampling class in the M-APP Region
Typical year
-Upper Midwest
U.S. Department of Agriculture
U.S. Forest Service
U.S. Geological Survey
Universal Transverse Mercator
13.1.2 Symbols
2As
2Bn
2Cn
2X
3A
3B
AC_BACL
AI3+
AL_AO
AL_CD
AL_CL
AL_CL2
AL_PYP
AW
C_TOT
Ca2+
CaCI2
CA_CL2
CAJDAC
CEC_CL
CECJDAC
cr
CLAY
C02
CON
Southern Blue Ridge subregion (NSS Pilot Survey)
Valley and Ridge subregion (NSS Pilot Survey)
Northern Appalachians subregion (NSS Pilot Survey)
Southern Appalachians subregion (NSS Pilot Survey)
Piedmont subregion (NSS Pilot Survey)
Mid-Atlantic Plain subregion (NSS Pilot Survey)
see Table 5-7
aluminum ion
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
total watershed area, in kilometers squared
carbon total
calcium ion
calcium chloride
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
chloride ion
see Table 5-7
carbon dioxide
mapped percent coniferous coverage.
357
-------
CONe
COS
COSI
effective percent coniferous coverage
see Table 5-7
see Table 5-7
FE_AO
FEJDD
FE_CL2
FE_PYP
FRAG
FS
FSI
H+
u-f
" total
H20
H2S04
ha
HC03-
H-DRY
H-WET
K
K+
K_CL
K_CL2
keq/ha
kg
km
KDAC
LTA-rbc
LTA-zbc
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
fragments > 2 mm diameter
see Table 5-7
see Table 5-7
hydrogen ion
total effective acidity (H+ + NH4+ -
water
sulfuric acid
hectare (2.47 acres or ten thousand square meters)
bicarbonate ion
annual hydrogen ion loading in dry deposition
annual hydrogen ion loading in wet deposition
hydraulic conductivity
potassium ion
see Table 5-7
see Table 5-7
kilooquivalent per hectare
kilogram
kilometer
see Table 5-7
sulfate mass transfer coefficient (m/yr)
long-term annual average, reduced dry base cation
long-term annual average, zero dry base cation
MA_MP_CM
MA_MP_UN
MA_SHEDS
MACMPNT
MAGIS01L
MATRAN
MATRANRC
/
-------
Mg2+
MG_CL
MG_CL2
MG_OAC
MOIST
MS
-magnesium ion
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
NJTOT
Na+
NA_CL
NA_CL2
NAOAC
NH
OH'
pC02
PH_002M
PH_01M
PH H20
- see Table 5-7
- sodium ion
— see Table 5-7
- see Table 5-7
- see Table 5-7
-ammonium ion
- nitrate ion
- hydroxide ion
-partial pressure of carbon dioxide
- see Table 5-7
- see Table 5-7
— see Table 5-7
R
r
R2
RTR
S_TOT
SAND
SBC_CI
sd
SiO2
SI_AO
SILT
SO4_0
SO4_2
SO4_4
SO4_8
SO4J6
S04_32
SO4_B2
SO4_EMX
S04 H2O
runoff (length/time)
correlation coefficient
coefficient of determination, the proportion of variability explained by a
regression model
lake retention time, in years
see Table 5-7
see Table 5-7
sum of base cations as measured in unbuffered 1N ammonium chloride
dry sulfur deposition (mass/length2/time)
silicon dioxide
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
see Table 5-7
half saturation constant
adsorption asymptote
sulfate, water extractable
359
-------
SO4_PO4
SO4_SLP
SO4_XIN
isotherms
S042'
— sulfate, phosphate extractable
- slope of sulfate adsorption isotherm at zero net adsorption
— zero net adsorption concentration for sulfate, determined from adsorption
-sulfate
- steady state sulfate concentration
— surface water sulfur (mass/length3)
- wet sulfur deposition (mass/length2/time)
- hydraulic residence time, in years
- hydrologic retention time, in years
VCOS
VFS
see Table 5-7
see Table 5-7
13.2 DEFINITIONS
ACCURACY - the difference between the approximate solution obtained using a numerical model and
the exact solution of the governing equations (or a known standard concentration), divided by the
exact solution (or known standard concentration).
ACID ANION - negatively charged ion that combines with hydrogen ion to form an acid.
ACID CATION - hydrogen ion or other metal that can hydrolyze water to produce hydrogen ions, e.g.
Al, Mn, Fe.
ACID CRYSTALLINE - in the Southern Blue Ridge Province, rocks or bedrock which, upon
weathering, form secondary phases including HIV clays.
ACID DEPOSITION SYSTEM (ADS) - a national database of precipitation amount and chemistry
maintained at Battelle-Pacific Northwest Laboratories.
ACID MINE DRAINAGE - runoff with high concentration of metals, sulfate, and acidity resulting from
the OXIDATION of sulfide minerals that have been exposed to air and water (usually from mining
activities).
ACID NEUTRALIZING CAPACITY - the total acid-combining capacity of a water sample determined
by titration with a strong acid to a preselected equivalence point pH: an integrated measure of the
ability of an aqueous solution to neutralize strong acid inputs. Acid neutralizing capacity includes
strong bases (e.g., hydroxide) as well as weak bases (e.g., borates, carbonates, dissociated organic
acids, alumino-hydroxyl complexes).
360
-------
ACIDIC DEPOSITION - rain, snow, or dry fallout containing high concentrations of sulfuric acid, nitric
acid, or hydrochloric acid, usually produced by atmospheric transformation of the by-products of fossil
fuel combustion (power plants, smelters, autos, etc.). Precipitation with a pH of less than 5.0 is
generally considered to be unnaturally acidic, i.e., altered by ANTHROPOGENIC activities.
ACIDIC EPISODE - an episode in a water body in which ACIDIFICATION of SURFACE WATER to an
ACID NEUTRALIZING CAPACITY less than or equal to 0 |ieq/L occurs.
ACIDIC LAKE OR STREAM - an aquatic system with an ACID NEUTRALIZING CAPACITY less than or
equal to 0 ^eq/L.
ACIDIFICATION - any temporary or permanent loss of ACID NEUTRALIZING CAPACITY in water or
BASE SATURATION in soil by natural or ANTHROPOGENIC processes.
ACIDIFIED - a natural water that has experienced any temporary or permanent loss of ACID
NEUTRALIZING CAPACITY or a soil that has experienced a reduction in BASE SATURATION.
ACTIVITY COEFFICIENTS - empirically derived coefficients used to transform concentration data to
salt or ion activities.
ADJUSTED R2 - the standard R2 of regression analysis, modified to balance increasing the R2 against
increasing the number of explanatory variables.
AFFORESTATION - the natural process through which non-forested lands become forested.
AGGRADING FORESTS - forests in which there is a net annual accumulation of biomass.
AGGREGATION - a method for statistically reducing a set of data to a single calculated or index value
for each parameter (e.g., a weighted average).
ALFISOLS - in Soil Taxonomy, the ORDER of mineral soils with an argillic horizon with at least 35
percent base saturation.
ALKALINITY - the titratable base of a sample containing hydroxide, carbonate, and bicarbonate ions,
i.e., the equivalents of acid required to neutralize the basic carbonate components.
ALKALINITY MAP CLASS - a geographic area defined by the expected ALKALINITY of SURFACE
WATERS (does not necessarily reflect measured alkalinity); used as a STRATIFICATION FACTOR in
ELS-I design.
ALUMINUM BUFFERING - a chemical process in which hydrogen ion activities are buffered by the
precipitation/dissolution of aluminum hydroxides.
ALUMINUM BUFFER RANGE - pH 4.2 - 2.8
361
-------
ANAEROBIC - without free oxygen (e.g., hypolirnnetic lake waters, sediments, or poorly drained soils).
ANALYTE - a chemical species that is measured in a water soil, or tissue sample.
ANALYTICAL CHARACTERIZATION - physical and chemical properties of water soil, or samples
measured in the laboratory.
ANALYTICAL DUPLICATE - a QUALITY CONTROL sample made by splitting a sample.
ANION - a negatively charged ion.
ANION CATION BALANCE - a method of assessing whether all CATIONS and ANIONS have been
accounted for and measured accurately; in an electrically neutral solution, such as water, the total
charge of positive ions (cations) equals the total charge of negative ions (anions).
ANION EXCHANGE/ADSORPTION - a reversible process occurring in soil in which ANIONS are
adsorbed and released.
ANTHROPOGENIC - of, relating to, derived from, or caused by human activities or actions.
APPARENT SOLUBILITY PRODUCT - an approximate form of an equilibrium constant calculated using
solution concentration data instead of activities.
AQUEOUS SPECIES - any dissolved ionic or nonionic chemical entity.
AQUIC - a moisture regime of soils in which a water table and reducing conditions occur near the
surface.
AQUIFERS - below-ground stratum capable of producing water as from wells or springs.
AQUO LIGAND - a water molecule held to Fe or Al in a clay edge or hydrous oxide by ligand exchange.
ARC - represents line features and borders of area features. One line feature may be made up of many
arcs. The arc is the line between two nodes.
ARC/INFO - a commercial geographic information system (GIS) software used to automate, manipulate,
analyze, and display geographic data in digital form.
ATTRIBUTE - the class, characteristics or other properties associated with a specific feature, area on a
map, lake or stream.
AVAILABLE TRANSECT - a transect identified to represent a map unit and listed for random selection.
362
-------
BASE CATION - a nonprotolytic CATION that does not affect ACID NEUTRALIZING CAPACITY;
consists principally of calcium, magnesium, sodium, and potassium.
BASE CATION EXCHANGE - the process by which BASE CATIONS (Ca2+, Mg2+, Na+, K+) are
adsorbed or released from negatively charged sites on soil particles from or to, respectively, soil
solutions. Such exchange processes are instrumental in determining pH of soil solutions.
BASE CATION SUPPLY - the pool of BASE CATIONS (Ca2+, Mg2+, Na+, K+) in a soil available for
exchange with hydrogen ions (H+). The base cation pool is determined by the CATION EXCHANGE
CAPACITY of the soil and the percentage of exchange sites occupied by BASE CATIONS.
BASE SATURATION - the percentage of total soil CATION EXCHANGE CAPACITY that is occupied by
exchangeable cations other than hydrogen and aluminum, i.e., the base cations Ca2+, Mg2+, Na+,
and K+.
BEDROCK - solid rock exposed at the surface of the earth or overlain by unconsolidated material.
BEDROCK GEOLOGY - the physical and chemical nature and composition of solid rock at or near the
earth's surface.
BEDROCK LITHOLOGY - see L1THOLOGY.
BEDROCK SENSITIVITY SCORES - a six point scale, developed for DDRP, designed to distinguish
the relative reactivities of different lithologies.
BEDROCK UNITS - the smallest homogenous entity depicted on a bedrock map.
BIAS - a systematic error in a method caused by artifacts or idiosyncracy of the measurement system.
BIOMASS - the quantity of particulate organic matter in units of weight or mass.
BIOMASS ACCRETION - net accumulation of plant mass in a growing, or aggrading, ecosystem; also
refers to net accumulation of an individual nutrient associated with accumulation of biomass.
BLOOM-GRIGAL MODEL - a numerical model used to investigate the evolution of soil exchange
characteristics under various H+ ion deposition SCENARIOS. The code is based on mass balance
consideration with empirical functions used to describe the pH-base saturation relationships.
BONFERRONI INEQUALITY - an inequality from probability theory that is used to carry out multiple
simultaneous statistical comparisons.
BOXPLOT - a graph of data with a box drawn from the 25th percentile to the 75th percentile; lines
extending from the box as far as the data extend to a distance of at most 1.5 times the
INTERQUARTILE RANGE, and more extreme observations marked individually.
363
-------
BUFFERING CAPACITY - the quantity of acid or base that can be added to a water sample with little
change in pH.
BULK DENSITY - the integrated density of a volume of soil, including solid matter, soil solutions,
voids, roots, etc.
Ca/AI EXCHANGE REACTION - the reaction describing the distribution of Ca and Al between the soil
exchange complex and the soil solution.
CALC1TE - a mineral with the formula CaCO3. A carbonate mineral.
CALIBRATION - process of checking, adjusting, or standardizing operating characteristics of
instruments and model appurtenances on a physical model or coefficients in a mathematical model
with empirical data of known quality. The process of evaluating the scale readings of an instrument
with a known standard in terms of the physical quantity to be measured.
CALIBRATION BLANKS - a zero-concentration QUALITY CONTROL standard that contains only the
matrix of the CALIBRATION standard.
CAPACITY FACTOR - a chemical property of a system defined as a function of the quantity or size of
that system.
CAPACITY-LIMITED PROCESS - A mechanism (e.g., sulfate adsorption or cation exchange) for which
the long-term ability to supply or consume cations or anions is constrained by the size of a watershed
pool or capacity (e.g., pool of exchangeable bases and sulfate adsorption capacity) rather than by
reaction kinetics.
CARBON-BONDED SULFUR - a reduced form of organic sulfur, characterized by C-S bonds.
CARBONIC ACID - a weak acid, H2CO3, formed by dissolution of carbon dioxide in water.
Dissociation of carbonic acid (to H+ and HCO3") and subsequent consumption of H+ by exchange or
weathering reactions generates ANC in the form of bicarbonate ions.
CATCHMENT - see WATERSHED.
CATION - a positively charged ion.
CATION DEPLETION - a process through which base cations on a soil exchange site are progres-
sively replaced by ACID CATIONS at rates higher than those expected during normal pedogenesis.
CATION EXCHANGE - a reversible process occurring in soil and/or sediment in which ACIDIC
CATIONS (e.g., hydrogen ions) are adsorbed and BASE CATIONS are released.
CATION EXCHANGE CAPACITY - the sum total of exchangeable cations that a soil can absorb.
364
-------
CATION (OR ANION) LEACHING - movement of cations (or anions) out of soil, in conjunction with
mobile anions in soil solution.
CATION RETENTION - the physical, biological, and geochemical processes by which cations in
watersheds are held, retained, or prevented from reaching receiving SURFACE WATERS.
CHRONIC ACIDIFICATION - see LONG-TERM ACIDIFICATION.
CIRCUMNEUTRAL - close to neutrality with respect to pH (pH = 7); in natural waters, pH 6 - 8.
CLAY - a soil separate consisting of particles with an equivalent diameter less than 0.002 mm; also a
soil textural class containing > 40 percent clay-sized material, < 45 percent sand and < 40 percent
silt.
CLAY MINERALS - any of a series of sheet silicate minerals formed in a soil or low-temperature
diagenetic environment.
CLOSED LAKES - a lake with a surface water inlet but no surface water outlet.
CLUSTER ANALYSIS - a multivariate classification technique for identifying similar (or dissimilar)
groups of observations.
COARSE PARTICLE DRY DEPOSITION - atmospheric DRY DEPOSITION of particles greater than 2
microns in effective diameter.
COLLINEAR - see MULTICOLLINEARITY.
COMBINATION BUFFER - land area surrounding a lake including area within a 40-foot elevation
contour, area within a linear buffer adjacent to perennial streams, and area around contiguous
wetlands.
COMPLEX - a map unit consisting of two or more dissimilar soil components or miscellaneous areas
occurring in a regularly repeating pattern.
COMPONENTS - see MAJOR COMPONENTS, MINOR COMPONENTS, and MAP UNIT
COMPOSITION.
CONSOCIATION - a map unit dominated by a single soil taxon (or miscellaneous area) and similar
soils.
CONTOUR LINE - a line connecting the points on the land surface that have the same elevation.
CONVERGENCE - state of tending to a unique solution. A given scheme is convergent if an
increasingly finer computational grid leads to a more accurate approximation of the unique solution.
Note that a numerical method may sometimes converge on a wrong solution.
365
-------
COVERAGE - a digital analog of a single map sheet; forms the basic unit of data storage in
ARC/INFO.
CUMULATIVE DISTRIBUTIVE FUNCTION - a function, F(x), such that for any reference value X, F(x)
is the estimated proportion of individuals (lakes, streams, estuaries, coastal waters) in the population
having a value x :< X.
DATABASE FILE - a collection of records that share the same format.
DEPOSITIONAL FLUXES - the mass transfer rate to the earth's surface of any of a number of
chemical species.
DEPTH TO BEDROCK - depth to solid, fixed, unweathered rock underlying soils.
DEPTH TO A SLOWLY PERMEABLE OR IMPERMEABLE LAYER - depth to a layer in soils or
underlying soils that restricts the downward flow of water (e.g., bedrock, dense till or fragipan).
DETECTION LIMIT QC CHECK SAMPLE - a QUALITY CONTROL sample that contains the ANALYTE
of interest at two to three times the contract required detection limit.
DIGITIZATION - the process of entering lines or points into a GEOGRAPHIC INFORMATION SYSTEM.
DIGITIZED COORDINATES - lines or points that have been entered into a GEOGRAPHIC
INFORMATION SYSTEM.
DISSIMILATORY REDUCTION - a process in which an oxidized chemical species (e.g., SO4 - S) is
utilized by an organism as an electron acceptor in the absence of free oxygen and released in a
reduced form (e.g., S2") rather than assimilated.
DISSOCIATION - separation of an acid into free H+ and the conjugate base of that acid (e.g.,
H2CO3 -> H+ + HCO3"), or separation of a base into a free hydroxyl and the conjugate acid of the
base (e.g., NH4OH -> NH4+ + OH").
DISSOLUTION RATES - the rate at which a mineral is transformed to aqueous species or secondary
minerals in an aqueous environment.
DISSOLVED ORGANIC CARBON - a measure of organic (nonorganic) fraction of carbon in a water
sample that is dissolved or unfilterable.
DOWNSTREAM REACH NODE - see LOWER NODE.
DRAINAGE - the frequency and duration of periods when the soil is free of saturation or partial
saturation and the depth to which saturation commonly occurs.
DRAINAGE BASIN - see WATERSHED.
366
-------
DRAINAGE CLASS - any of the seven classes that characterize the frequency and duration of soil
saturation.
DRAINAGE LAKE - a lake with SURFACE WATER outlet(s) or with both inlets and outlets.
DRY DEPOSITION - for the purposes of DDRP analysis, atmospheric deposition of materials to
watersheds in any form other than rain or snow.
DRY DEPOSITION VELOCITY - an effective velocity used with airborne concentrations to compute dry
depositional flux of materials to surfaces or watersheds.
ELECTRON ACCEPTOR - an oxidized (or at least partially oxidized) chemical species capable of
undergoing a reduction reaction by addition of an electron.
ELS PHASE I LAKES - the population of lakes sampled during phase I of the Eastern Lake Survey of
the EPA's National Surface Water Survey.
EMPIRICAL MODEL - representation of a real system by a mathematical description based on
experimental data rather than on general physical laws.
ENTISOLS - in Soil Taxonomy, the ORDER of mineral soils with no or very poorly developed genetic
horizons.
EPISODE - a short-term change in stream pH and ACID NEUTRALIZING CAPACITY during storm flows
or snowmelt runoff.
EQUIVALENT - unit of ionic concentration; the quantity of a substance that either gains or loses one
mole of protons or electrons.
ESTER SULFATE - an oxidized form of sulfur in soil organic matter, characterized by C-O-SO3 or N-O-
SO3 linkages.
EVAPOTRANSPIRATION (%ET) - the amount or proportion of precipitation that is returned to the air
through direct evaporation or by transpiration of vegetation.
EXCHANGE POOL - the reservoir of BASE CATIONS in soils available to participate in exchange
reactions.
EXTENSIVE PARAMETERS - variables that depend on the size (extent) of the system.
EXCHANGE REACTIONS - any of a number of reactions that describe the partitioning of two chemical
species between a solution and soil exchange complex.
FELDSPARS - a group of tectosilicate minerals that are the most abundant group in the earth's crust.
367
-------
FIELD REVIEW - a review of soil surveys made in the field by supervisory soil scientists to help field
soil scientists maintain standards that are both adequate for the objectives of the survey and
consistent with those of other surveys. Samples of the fieldwork are examined for soil identification,
placement of boundaries, and map detail in relation to survey objectives.
FINE PARTICLE: DRY DEPOSITION - atmospheric DRY DEPOSITION of particles of size less than 2
microns in effective diameter.
FIRST-ORDER REACTION - a chemical reaction, the rate of which is proportional to the concentration
of the limiting reactant.
FOREST COVER TYPE - a descriptive classification of forest land based on present occupancy of an
area by tree species. (The term 'Vegetation" implies total forest community, whereas the focus here is
on trees defining type. Whenever the term "vegetation" is used in this report it should be construed as
FOREST COVER TYPE.)
FREUNDLICH ISOTHERM - an exponential adsorption isotherm of the form E0 = aCb, where: Ec =
concentration of adsorbed species (per unit mass adsorbent), C = dissolved concentration of species
being adsorbed, and a and b are derived coefficients.
FULVIC ACID - a family of naturally-occurring weak organic acids found in soils and surface waters;
fulvic acids are operationally defined as the acid-soluble (pH = 1.0) fraction of an alkali-soluble soil
extract; pK is roughly 3.5.
GAINES THOMAS FORMULATION - a formulation used to describe exchange processes.
GAPON - a formulation used to describe exchange processes.
GENERIC BEDROCK TYPE - see GENERIC ROCK TYPE.
GENERIC ROCK TYPE - a general classification of different BEDROCK UNITS into groups according
to the primary LITHOLOGY.
GEOGRAPHIC INFORMATION SYSTEM (GIS) - a computerized system designed to store, process,
and analyze data.
GEOLOGY - see BEDROCK GEOLOGY.
GEOMORPHIC POSITION - the relative location in the landscape described by hillslope elements
(cross section view) and slope components (plane view), e.g., sideslope footslope.
GIBBSITE - a mineral with the chemical formula AI(OH)3.
368
-------
GIS BUFFERS - land area surrounding a lake, stream, or wetland, delineated using a GIS. See
COMBINATION BUFFER and LINEAR BUFFER.
GLACIAL TILL - see TILL
GLACIOFLUV1AL - a material that has been deposited by glaciers and sorted by meltwater.
GLEY SOIL - a soil developed under conditions of poor drainage, characterized by oxygen depletion
and reduction of iron and other metals (Mn), resulting in gray colors and mottles.
GRAN ANALYSIS - a mathematical procedure used to determine the equivalence points of a
TITRATION CURVE for acid and base neutralizing capacity.
GROUNDWATER - water in the part of the ground that is completely saturated.
HEURISTIC MODEL - representation of a real system by a mathematical description based on
reasoned, but unproven argument, intended for use as an aid to studying and exploratory analysis of
the system being modelled.
HINDCAST - to estimate some prior event or condition as a result of a rational study and analysis of
available pertinent current and historical data.
HISTOSOLS - in Soil Taxonomy, the ORDER of soils formed from organic PARENT MATERIAL.
HORIZON - a horizontal layer of soil with distinct physical and/or chemical characteristics. Genetic
horizons are the result of soil-forming process.
HYDRAULIC HEAD - hydrostatic pressure created by a difference in height of water columns in
different portions of a connected aquifer.
HYDRAULIC RESIDENCE TIME - a measure of the average amount of time water is retained in a lake
basin. It can be defined on the basis of inflow/lake volume, represented as RT, or on the basis of
outflow/lake volume and represented as Tw. The two definitions yield similar values for fast flushing
lakes, but diverge substantially for long residence time SEEPAGE LAKES.
HYDROLOGIC FLOW PATHS - the distribution and circulation of water deposited by precipitation on
the surface of the land, in the soil, and underlying rocks within a WATERSHED.
HYDROLOGIC RETENTION TIME - see HYDRAULIC RESIDENCE TIME.
HYDROUS OXIDE - a collective term referring to any of a group of amorphous or crystalline species
of iron or aluminum that are partially or fully hydrated (e.g., MO(OH), M(OH)3).
369
-------
IMMOBILIZATION REACTION - conversion of an inorganic form of a nutrient (especially S or N) to
organic matter.
IMPOUNDMENT - a man-made lake created by construction of a dam; also applied to natural lakes
whose level is controlled by a human-made spillway.
INCEPTISOLS - in Soil Taxonomy, the ORDER of soils with at least one diagnostic horizon, but with
no horizon strongly enough developed to place them in another ORDER.
INCLUSIONS - see MINOR COMPONENTS.
INDEX OF CONTACT TIME - the theoretical maximum potential of contact between runoff and the
soil matrix. The index is calculated by dividing the soil water flow rate (obtained using Darcy's law) by
average annual runoff.
INDEX SAMPLE - in NE lakes, one sample per lake, used to represent chemical conditions on that
lake. In streams, any sample (or the average of one to three samples) collected at a stream NODE
during the SPRING BASEFLOW INDEX PERIOD, used to represent chemical conditions in the stream.
INFO - a database management system that stores, maintains, manipulates, and reports information
associated with geographic features in ARC/INFO.
INITIAL CONDITIONS - given values of DEPENDENT VARIABLES or relationship between dependent
and independent variables at the time of start-up of the computation in a mathematical model.
IN-LAKE SULFUR RETENTION - net retention of sulfur within a lake, occurring principally by
reduction within sediments.
INTENSITY FACTOR - a variable with properties defined by concentration or a surface and/or in
solution, and therefore independent of the quantity or size of the system.
INTENSIVE PARAMETERS - variables whose values are independent of the size or extent of the
system, e.g., temperature and pH.
INTERQUARTILE RANGE - the difference between the 75th and the 25th percentiles.
IONIC STRENGTH - a measure of the interionic effect resulting from the electrical attraction and
repulsion between various ions. In very dilute solutions, ions behave independently of each other and
the ionic strength can be calculated from the measured concentrations of ANIONS and CATIONS
present in the solution.
370
-------
ISOTHERM - a linear or nonlinear function describing partitioning of an absorbent between solid and
sorbed phases. Such isotherms were originally used to characterize (nearly) ideal processes (e.g., the
Langmuir equation was developed to describe adsorption of a gas by a solid), but are often
empirically defined for adsorption of anions or organic compounds on soils because they provide a
convenient shorthand to describe partitioning.
KINETIC MODELS - any of a family of numerical models that use kinetic considerations as the
unifying principle in describing natural processes.
KRIGING - a technique for spatial interpolation.
LABEL - represents point features or is used to assign identification numbers to POLYGONS.
LAKE TYPE - a classification of lakes based on the presence or absence of inlets, outlets, and dams
as represented on LARGE-SCALE MAPS.
LAND COVER - see FOREST COVER TYPE.
LAND USE - the dominant use of an area of land (e.g., crop land).
LANDFORM SEGMENT - a small part of the local landform that is uniquely related to landscape
processes.
LANGMUIR ISOTHERM - a hyperbolic adsorption isotherm (used in this project for sulfate) of the
form Ec = (B1 * C)/(B2 + C), where: Ec = net adsorbed sulfate, C = dissolved sulfate, and B1 and B2
are empirically derived coefficients. When appropriate, the isotherm can be "extended" by addition of
a third coefficient to describe a non-zero Y-intercept.
LARGE-SCALE MAPS - 1:24,000, 1:25,000, or 1:62,500 scale U.S. Geological Survey topographical
maps.
LEACHING - the transport of a solute from the soil in the soil solution.
LEVERAGE POINT - a data point that strongly influences the parameter estimates in a regression.
LIGAND EXCHANGE - a mechanism of bond formation between an oxyanion and a soil mineral
bearing hydroxyl groups. The exchange involves formation of inner sphere complexes of anions to
Lewis acid sites, following replacement of water from the Lewis acid site by the oxyanion.
LIMESTONE - a rock type consisting primarily of CALCITE.
LINEAR BUFFER - land area within a set distance of a lake or stream.
371
-------
LITHOLOGY - the physical characteristics of a rock or mapped BEDROCK UNIT. Generally relates to
mode of formation, mineralogy, and texture.
LITTERFALL - fresh organic detritus, usually leaves, needles, twigs, etc., that compose the bulk of the
forest floor.
LOCAL LANDFORM - a subdivision of the regional landform that is the result of localized landscape
processes.
LONG-TERM ACIDIFICATION - a long-term partial or complete loss of ACID NEUTRALIZING CAPACITY
from a lake or stream.
LONG-TERM ANNUAL AVERAGE DEPOSITION (LTA) - a dataset of atmospheric deposition
representing atmospheric deposition during the early-to-mid 1980s for the purposes of the DDRP.
LOWER NODE - the downstream NODE of a STREAM REACH.
MAJOR LAND RESOURCE AREA - a geographic area characterized by a particular pattern of soils,
climate, water resources, and LAND USE.
MAJOR COMPONENTS - soil components or miscellaneous areas that are identified in the name of a
map unit.
MAP COMPILATION - the process of checking and measuring soil map unit data.
MAPPING PROTOCOLS - instructions that guide the field mapping and provide for quality control.
MAP SYMBOL - a symbol used on a map to identify map units.
MAP'PING TASK LEADER - the person responsible for field mapping activities.
MAP UNIT - see SOIL MAP UNIT.
MAP UNIT COMPOSITION - the relative proportion (expressed in percent) of all soil components and
miscellaneous areas in a map unit.
MAP UNIT COMPOSITION FILE - a DATABASE FILE that contains ail components and their relative
proportion for each map unit in the survey area (components are identified by an assigned code, i.e.,
SCODE).
MAP UNIT CORRELATION - see SOIL CORRELATION.
MAP UNIT DELINEATION - an area on a map uniquely identified with a symbol. A delineation of a soil
map has the same major components as identified and named in the map unit.
372
-------
MAP UNIT NAME - the title of a map unit identified by the major soil components or miscellaneous
areas followed by appropriate phase terms.
MASS ACTION MODELS - any of a family of numerical models that use equilibrium-based principles as
the central unifying theme.
MASS BALANCE MODELS - any of a family of numerical models that use conservation-of-mass
principles as the central unifying theme.
MASS TRANSFER COEFFICIENTS - a removal or rate constant used in models of in-lake alkalinity
generation (and elsewhere) to quantify the average removal rate of a reactant from solution. Specific
reference in this project is transfer from solution to sediment by all processes, including sedimentation
and diffusion. In many systems, the mass transfer coefficient for sulfur is essentially a diffusion constant
for sulfate across the water-sediment interface; for nitrate a biological uptake/sedimentation rate.
MASTER HORIZONS - the most coarsely based delineations within a pedon. Usually, A/E horizons
denote zones of net mass depletion, B horizons are zones of net accumulation, and C horizons indicate
minimal pedogenic evolution.
MATRIX SPIKE - a QUALITY CONTROL sample made by adding known quantity of an ANALYTE to a
sample aliquot.
MAX - the maximum sensitivity code observed on a WATERSHED.
MEAN - the weighted average of sensitivity codes for a WATERSHED.
MEDIAN (M) - the value of x such that the cumulative distribution function F(x) = 0.5; the 50th
percentile.
MID-APPALACHIAN REGION - one of the three geographic regions considered by the, DDRP,
consisting of upland areas (subregions 2Bn and 2Cn) of the Mid-Atlantic region (MD, PA, VA, WV)
defined by the National Stream Survey.
MINERAL WEATHERING - dissolution of rocks and minerals by erosive forces.
MINERALIZATION - microbially-mediated conversion of nutrients from an organically bound (especially
N and S) to an inorganic form.
MINOR COMPONENTS - soil components or miscellaneous areas that are not identified in the name of
the map unit. Many areas of these components are too small to be delineated separately.
MISCELLANEOUS AREA - land areas that have no soil and thus support little or no vegetation without
major reclamation. Rock outcrop is an example.
373
-------
MISCELLANEOUS LAND AREAS - see MISCELLANEOUS AREA.
MOBILE ANION - an anion that remains in solution and passes through a soil without significant delays
due to biological or chemical processes; also a model or paradigm for cation leaching from soils, based
on the premise that the rate of cation leaching from a soil is controlled by the sum of mobile anions
(which are regulated by a suite of more-or-!ess independent processes).
MONTE CARLO METHOD - technique of STOCHASTIC sampling or selection of random numbers to
generate synthetic data.
MOTTLING - spots or blotches of different color in a soil, including gray to black blotches in poorly
drained soils due to presence of reduced iron and other metals.
NITROGEN TRANSFORMATION - biochemical processes through which nitrogen deposited in an
environment is converted to other forms.
NODE - the points identifying either an upstream or downstream end of a REACH.
NONPARAMETRIC - referring to a statistical procedure that does not make the classical distributional
assumptions.
NON-SILICATE IRON AND ALUMINUM - soil iron and/or aluminum occurring in the soil as an
amorphous or a (hydrous) oxide phase rather than as an ion incorporated within a silicate mineral
lattice.
OFFICIAL SOIL SERIES DESCRIPTION - a record of the definitions of a soil series and other relevant
Information about each series. These definitions are the framework within which most of the detailed
information about soils of the United States is identified with soils at specific places. These definitions
also provide the principal medium through which detailed information about the soil and its behavior at
one place is projected to similar soils at other places.
ORDER - in Soil Taxonomy, the highest level of classification, e.g., SPODOSOLS.
ORGANIC ACID - organic compound possessing an acidic functional group; includes fulvic and humic
acids.
ORGANIC ANION - an organic molecule with a negative net ionic charge.
ORGANIC "BLOCKING" - a reduction in the sulfate (or other anion) adsorption capacity of a soil
resulting from preferential sorption of organic acids by the soil.
ORGANIC HORIZONS - any identifiable soil horizon containing in excess of 20 percent organic matter
by weight.
374
-------
OUTLIER - observation not typical of the population from which the sample is drawn.
OXIDATION - loss of electrons by a chemical species, changing it from a lower to a higher oxidation
state (e.g., Fe2+ to Fe3+ or S'2 to S+6, with intermediates).
PARAMETER - (1) a characteristic factor that remains at a constant value during the analysis, or (2) a
quantity that describes a statistical population attribute.
PARENT MATERIAL - the material from which soils were formed.
PARTIAL PRESSURE - the pressure of a gaseous sample that is attributable to one particular
component.
PEDON - the smallest block of soil that contains all the characteristics of a soil (usually about 1 m2); a
soil individual.
PERCENT COARSE FRAGMENTS - the percentage of soil, by volume, that is composed of rock
fragments unable to pass through a 2-mm sieve.
PERMEABILITY - the ease with which gases, liquids, or plant roots penetrate or pass through a bulk
mass of soil or a layer of soil.
pH - the negative logarithm of the hydrogen ion activity. The pH scale runs from 0 (most acidic) to 14
(most alkaline); a difference of 1 pH unit indicates a tenfold change in hydrogen activity.
POLYGON - represents area features.
PRECISION - a measure of the capacity of a method to provide reproducible measurements of a
particular ANALYTE (often represented by variance).
PRIMARY MINERAL WEATHERING - the natural process by which thermodynamically unstable
minerals are converted to more stable phases under earth surface conditions.
PRINCIPAL COMPONENT ANALYSIS - a statistical analysis concerned with explaining the variance-
covariance structure through the use of PRINCIPAL COMPONENTS.
PRINCIPAL COMPONENTS - particular linear combinations of the original data, which geometrically
represent a new coordinate system with axes in the directions of maximum variability.
PROBABILITY SAMPLE - a sample in which each unit has a known probability of being selected.
QC CHECK SAMPLE - a QUALITY CONTROL sample that contains the ANALYTE of interest at a
concentration in the mid-calibration range.
375
-------
QUALITY ASSURANCE - steps taken to ensure that a study is adequately planned and implemented
to provide data of known quality, and that adequate information is provided to determine the quality of
the database resulting from the study.
QUALITY CONTROL - steps taken during a study to ensure that data quality meets the minimum
standards established by the quality assurance plan.
QUARTILE - any of three values (Q1, Q2, Q3) that divide a population into four equal classes, each
containing one-fourth of the population.
QUINTILE - any of the four values (Q1 , Q2 , Q3 , Q4) that divide a population into five equal classes,
each representing 20 percent of the.population; used to provide additional values to compare
characteristics among populations of lakes and streams.
RATE-LIMITED REACTION - a process (e.g., mineral weathering) for which the long-term ability to
supply reaction products (e.g., base cations) is constrained by reaction or transport kinetics.
RCC TRANSECTS - transects conducted by the Regional Coordinator/Correlator (RCC).
REACH - segments of the stream network represented as blue lines on 1:250,000-scale U.S.
Geological Survey maps. Each reach (segment) is defined as the length of stream between two blue-
line confluences. In the NSS-I, stream reaches were the sampling unit.
REACTION ORDER - the relationship between the rate of a chemical reaction and the concentration
of a reaction substrate, defined by the value of the exponent of that substrate.
REAGENT BLANK - a QUALITY CONTROL sample that contains all the reagents used and in the
same quantities used in preparing a soil sample for analysis.
REDUCTION/OXIDATION - chemical reaction in which substances gain or lose electrons.
REGION - a major area of the conterminous United States where a substantial number of streams with
ALKALINITY less than 400 jieq/L can be found.
REGIONAL LANDFORM - physiographic areas that reflect a major land-shaping process over a long
period of time.
REGIONAL SOILS LEGEND - a correlated and controlled legend for an entire region (see SOIL
IDENTIFICATION LEGEND).
RELMAP - a source-receptor model designed to estimate dry deposition of sulfur; not used directly in
the DDRP.
376
-------
REPORTS - relative to GIS activities, a format designed by the user for printing out information
containing the data files.
RESERVOIR - a body of water collected and stored for future use in a natural or artificial lake.
RESIDUAL - in regressions, the difference between the observed dependent variable and the value
predicted from the regression fit.
REUSS MODEL - a numerical model used to describe exchange processes in a soil environment.
RIPARIAN - a zone bounding and directly influenced by SURFACE WATERS.
ROBUST - a statistical procedure that is insensitive to the effect of OUTLIERS.
ROUTINE TRANSECTS - transects conducted by field soil scientists responsible for the mapping.
SALT EFFECT - the process by which hydrogen ions are displaced for the soil exchange complex by
BASE CATIONS (from neutral salts). The result is a short-term increase in the acidity of associated
water.
SAMPLING CLASS - see SOIL SAMPLING CLASS.
SAMPLING CLASS CODE - a three-character code assigned to each SOIL SAMPLING CLASS.
SAMPLING CLASS COMPOSITION - the relative proportion of sampling classes in a map unit.
SAND - a soil separate between 0.05 and 2.0 mm in diameter; also a soil texture class containing at
least 85 percent sand, and whose percentage of silt, plus 1.5 times the percent clay, does not exceed
15.
SATURATION INDEX - the ratio of the ion activity product (of dissolved ions) to the solubility product
for a solid phase; if the saturation index (SI) exceeds 1.0, the solution is supersaturated with respect to
that phase; if SI = 1.0, the solution is at equilibrium, if SI < 1.0, the solution is undersaturated with
respect to that solid phase.
SCENARIO - one possible deposition sequence following implementation of a control or mitigation
strategy and the subsequent effects associated with this deposition sequence.
SECONDARY MINERALS - any inorganic mineral phase formed by transformation of another mineral or
by precipitation from an aqueous phase.
SEEPAGE LAKE - a lake with no permanent SURFACE WATER inlets or outlets.
377
-------
SELECTIVITY COEFFICIENT - the apparent constant used to describe the partitioning of species in
an exchange reaction.
SENSITIVITY ANALYSIS - test of a model in which the value of a single variable or parameter is
changed, and the impact of this change on the DEPENDENT VARIABLE is observed.
SENSITIVITY CODES - see BEDROCK SENSITIVITY SCORES.
SIGNIFICANCE LEVEL - the conditional probability that a statistical test will lead to rejection of the
null hypothesis, given that the null hypothesis is true.
SILICA - the dissolved form of silicon dioxide (SiO2).
SILT - a soil separate consisting of particles between 0.05 and 0.002 mm in equivalent diameter; also
a soil texture class containing at least 80 percent silt and < 12 percent clay.
SILVICULTURAL PRACTICES - forest management practices to increase wood yields: thinning,
pruning, fertilization, spraying with herbicides/insecticides, and irrigating.
SIMULATION - replication of the prototype using a model.
SKELETAL SOILS - soils with at least 35 percent rock fragments in the control section.
SLOPE PHASE - the slope gradient of a soil map unit or taxonomic unit expressed in percent.
SLOPE SHAPE ACROSS - shape of the surface parallel to the contours of the landscape (e.g.
concave, convex, plane).
SLOPE SHAPE DOWN - shape of the land surface at right angles to the contours of the landscape.
SMALL-SCALE MAP - 1:250,000-scale U.S. Geological Survey map.
SOIL - unconsolidated material on the surface of the earth that serves as a natural medium for the
growth of plants.
SOIL ACIDIFICATION - a process through which BASE CATIONS are removed from the soil and are
replaced by ACID CATIONS.
SOIL BUFFERING CAPACITY - the capacity of a soil to resist changes in pH with the addition of
acids to the sysrtem.
SOIL COMPONENT CODE - four-character code assigned to each soil or miscellaneous area
component of map units in a survey area. Codes were used to link data files.
378
-------
SOIL COMPONENTS FILE - a computer data file that contains all the soil and miscellaneous area
components in a survey area and identified with a code (i.e., SCODE).
SOIL CORRELATION - the process of maintaining consistency in naming, classifying, and interpreting
kinds of soils and of the units delineated on maps.
SOIL EXCHANGE COMPLEX - all components of a soil that contribute to the absence of exchange
properties of that soil.
SOIL FAMILY - next to the lowest category in Soil Taxonomy in which classes are separated mainly on
particle size, temperature, and mineralogy.
SOIL IDENTIFICATION LEGEND - a map legend that lists the symbols used to identify SOIL MAP
UNITS and the names of the map units.
SOIL LEGEND - see SOIL IDENTIFICATION LEGEND.
SOIL MAP UNIT - a collection of areas defined and named in terms of their soil components or
miscellaneous areas or both. Each map unit differs in some respect from all others in a survey area and
is uniquely identified on a soil map.
SOIL SAMPLING CLASS - an arbitrary grouping of soils either known or expected to have similar
physical and/or chemical effects on drainage waters with respect to effects of acidic deposition.
SOIL SERIES - the most homogenous category in the taxonomy used in the United States. A group of
soils that have horizons similar in arrangement and in differentiating characteristics.
SOIL SOLUTIONS - those aqueous solutions in contact with soils.
SOIL TAXONOMIC CLASS - the soil members within limits of ranges set by Soil Taxonomy. Taxonomic
units are members of the taxonomic class.
SOIL TAXONOMIC UNIT - a kind of soil described in terms Of ranges in soil properties of the
polypedons referenced by the taxonomic unit in the survey area.
SOIL TEXTURE - the relative proportion by weight, of the several soil particle size classes finer than 2
mm in equivalent diameter (e.g., sandy loam).
SOIL TEXTURE MODIFIER - suitable adjectives added to soil texture classes when rock fragments
exceed about 15 percent by volume, for example, gravelly loam. The terms 'Very" and "extremely" are
used when rock fragments exceed about 35 and 60 percent by volume, respectively.
SOIL TRANSECT - a distance on the surface of the earth represented by a line on a map. Transects
can be straight, dogleg, or zigzag.
379
-------
SOLID PHASE IEXCHANGERS - those components of soils, primarily organic matter, clay minerals
and mineral oxides, that serve as the sites for exchange reactions.
SOLUM - soil layers that are affected by soil formation.
SPECIATION MODEL - a numerical model used to describe the distribution of aqueous species
among various possible complexes and ion pairs; usually for the purpose of estimating single ion
activities.
SPECIFIC ADSORPTION - adsorption of sulfate by ligand exchange, often involving exchange of two
ligands and formation of a bridged (M-O-SO2-O-M) structure.
SPODIC HORIZONS - a soil hon'zon in which iron oxides, aluminum oxides, and organic matter have
accumulated from higher horizons.
SPODOSOLS - in Soil Taxonomy, the ORDER of mineral soils with well-developed SPODIC
HORIZONS.
SPRING BASEF:LOW INDEX PERIOD - a period of the year when streams are expected to exhibit
chemical characteristics most closely linked to ACIDIC DEPOSITION. The time period between
snowmelt and leafout (March 15 to May 15 in the NSS-I) when NSS-I stream reaches were visited
coinciding with expected periods of highest geochemical and assessment interest (i.e., low seasonal
pH and sensitive life stages of biota).
STABILITY (NUMERICAL OR COMPUTATIONAL) - ability of a scheme to control the propagation or
growth of small perturbations introduced in the calculations. A scheme is unstable if it allows the
growth of error so that it eventually obliterates the solution.
STANDARD DEVIATION - the square root of the variance of a given statistic.
STEADY-STATE - the condition that occurs when a property (e.g., mass, volume, concentration) of a
system does not change with time. This condition requires that sources and sinks of the property are
in balance (e.g., inputs equal outputs; production equals consumption).
STEPWISE REGRESSION - a statistical model-selection technique in which variables are added to the
model one at a time using F statistics as criteria, and then previously included variables are examined
for removal from the model.
STOCHASTIC - an occurrence that is probabilistic in nature.
STRATIFICATION FACTOR - factor used to define strata prior to lake selection; the factors used in
the ELS-I were region, subregion, and alkalinity map class.
STRATUM - in the NSS-I, sampling strata were designated as REGION and expected alkalinity map
class (greater than or less than 50 fieq/L).
380
-------
STREAM REACH - see REACH.
STREAM SAMPLING NODE - the points identifying either an upstream or downstream end of a
REACH. In the NSS-I, the points above or below the confluence of two streams.
SUBGROUP - in Soil Taxonomy, the level of classification just below GREAT GROUP, e.g., Typic
Haplorthods.
SUBORDER - in Soil Taxonomy, the level of classification just below ORDER, e.g., Orthods.
SUBREGIONS - areas within each region that are similar in water quality, physiography, vegetation,
climate, and soil; used as a STRATIFICATION FACTOR in ELS I design.
SUBTENDING - of or pertaining to subjacent or downslope landscape positions.
SULFATE ADSORPTION - reversible retention of sulfate on the surface of positively-charged surfaces
of soils, by either electrostatic attraction or by ligand exchange.
SULFATE ISOTHERM - a mathematical relationship between the amounts of sulfate absorbed on soil
and the concentrations of sulfate in solution equilibrated with that soil.
SULFATE MINERALS - a mineral containing sulfur in the +6 oxidation state (SO4); sulfate minerals
are typically formed as evaporites; gypsum and anhydrite are common sulfate minerals.
SULFATE MOBILITY - unrestricted movement (in the soil solution) of sulfate through the soil, with
subsequent release in soil leachate; also the extent to which sulfur supplied to a watershed from
external sources (i.e., acidic deposition) passes though watershed soils rather than being retained by
adsorption or some other process.
SULFATE RETENTION - the physical, biological, and geochemical processes by which sulfate in
WATERSHEDS is held, retained, or prevented from reaching receiving SURFACE WATERS.
SULFIDE - an ion consisting of reduced sulfur (S2") or a compound containing sulfide, e.g., hydrogen
sulfide (H2S) or the iron sulfide pyrite (FeS2).
SULFIDE OXIDATION - chemical reaction in which a sulfide loses electrons and assumes a higher
oxidation state; sulfate is the completely oxidized end product.
SULFIT1C - containing sulfide minerals, usually pyrite.
SULFUR INPUT/OUTPUT BUDGET - an approach to describing sulfur mobility in a watershed by
comparing fluxes of sulfur to and from the watershed (as the difference between input and output or
as a ratio).
381
-------
SURFACE WATER - streams and lakes.
SURFACE WATER RUNOFF - precipitation that flows overland to reach SURFACE WATERS.
SURFiCIAL GEOLOGY - characteristics of the earth's surface, especially consisting of unconsolidated
residual, colluvial, or glacial deposits lying on the BEDROCK.
SYNOPTIC - relating to or displaying conditions as they exist at a point in time over a broad area.
SYSTEMATIC ERROR - a consistent error introduced in the measuring process. Such error commonly
results in biased estimations.
TARGET POPULATION - a subset of a population explicitly defined by a given set of exclusion criteria
to which inferences are to be drawn from the sample attributes.
THERMODYNAMIC CONSTANTS - an empirically derived constant used to describe the relative
distribution of chemical species in a specified reaction when at equilibrium.
THROUGHFALL - precipitation that has interacted with a forest canopy, the chemistry of which is thus
modified from that of incident precipitation due to wash off of dry-deposited material and leaf exudates
as well as by Ion exchange and uptake by leaf surfaces.
TILL - unstratified material deposited by glaciers.
TITRATION CURVES - a loci of points describing some solution property, usually pH, as a function of
the sequential addition of a strong acid (or base) to the system.
TOPMODEL - topographically based, variable source area hydrologic model.
TOPOGRAPHIC MAP - a map showing contours of surface elevation.
TRANSECT - see SOIL TRANSECT.
TRANSECTING - a field activity involving the collection of data at points along a designated line (see
TRANSECT POINTS).
TRANSECT POINTS - locations along a TRANSECT where data are collected.
TRANSECT SEGMENT UNION - all transect stops in the same map unit on a WATERSHED.
TRANSECT STOPS - see TRANSECT POINTS.
TRANSFORMATION ERROR - calculates the residual mean square error of the digitized TIC locations
and the existing TICs.
382
-------
TRAVERSING - a field activity that involves observation at uncontrolled representative locations in the
landscape.
TYPICAL YEAR (TY) DEPOSITION DATA - a dataset of atmospheric deposition developed within the
DDRP for specific use with the integrated watershed models.
UNCERTAINTY ANALYSIS - the process of partitioning modelling error or uncertainty to four sources:
intrinsic natural variability, prior assumptions/knowledge, model identification, and prediction error.
UNIVERSAL TRANSVERSE MERCATOR (UTM) PROJECTION - a standard map projection used by
the U.S. Geological Survey.
UPPER NODE - the upstream NODE of a STREAM REACH.
UPSTREAM REACH NODE - see UPPER NODE.
UTM COORDINATES - lines or points as represented in a UNIVERSAL TRANSVERSE MERCATOR
PROJECTION.
VALIDATION - comparison of model results with a set of prototype data not used for verification.
Comparison includes the following: (1) using a dataset very similar to the verification data to
determine the validity of the model under conditions for which it was designed; (2) using a dataset
quite different from the verification data to determine the validity of the model under conditions for
which it was not designed but could possibly be used; and (3) using post-construction prototype data
to determine the validity of the predictions based on model results.
VANSELOW EXCHANGE FORMULATION - a formulation used to describe soil exchange reactions.
VARIABLE - a quantity that may assume any one of a set of values during the analysis.
VARIABLE SOURCE AREA - A topographically convergent, low transmissivity area within a watershed
that tends to produce saturation excess overland flow during storm runoff periods.
VEGETATION - see FOREST COVER TYPE.
VERIFICATION - check of the behavior of an adjusted model against a set of prototype conditions.
WATERSHED - the geographic area from which SURFACE WATER drains into a particular lake or
point along a stream.
WATERSHED STEADY STATE - a condition in which inputs of a constituent to a WATERSHED equal
outputs.
383
-------
WATERSHED SULFUR RETENTION - retention of sulfur by any of a number of mechanisms within a
WATERSHED.
WEATHERED BEDROCK - soft or partly consolidated BEDROCK that can be dug with a spade.
WEATHERING - physical and chemical changes produced in rocks at or near the earth's surface by
atmospheric agents with essentially no transport of the altered materials.
WEIGHT - the inverse of a sample's inclusion probability; each sample site represents this number of
sites in the TARGET POPULATION.
WET DEPOSITION - for the purposes of the DDRP, atmospheric deposition of materials via rain or
snow.
WETLAND - an area, generally with hydric soils, that is saturated, flooded, or ponded long enough
during the growing season to develop anaerobic conditions in the upper soil horizons and that is
capable of supporting the growth of hydrophitic vegetation.
ZERO-ORDER REACTION - a chemical reaction, the rate of which is independent of reactant
concentration.
384
------- |